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/9605091 | 1 | 9605 | 1996-05-15T19:11:33 | Theory of DDT in Unconfined Flames | [
"astro-ph"
] | This paper outlines a theoretical approach for predicting the onset of detonation in unconfined turbulent flames which is relevant both to problems of terrestrial combustion and to thermonuclear burning in Type Ia supernovae. Two basic assumuptions are made: 1) the gradient mechanism is the inherent mechanism that leads to DDT in unconfined conditions, and 2) the sole mechanism for preparing the gradient in induction time is by turbulent mixing and local flame quenching. The criterion for DDT is derived in terms of the one-dimensional detonation wave thickness, the laminar flame speed, and the laminar flame thickness in the reactive gas. This approach gives a lower-bound criterion for DDT for conditions where shock preheating, wall effects, and interactions with obstacles are absent. Regions in parameter space where unconfined DDT can and cannot occur are determined. A subsequent paper will address these issues specifically in the astrophysical context. | astro-ph | astro-ph |
A Theory of DDT in Unconfined Flames
Alexei M. Khokhlov,* Elaine S. Oran,** J. Craig Wheeler*
* Department of Astronomy
University of Texas at Austin
Austin, TX 78712
** Laboratory for Computational Physics
Naval Research Laboratory
Washington, DC 20375
Corresponding Author:
Dr. Alexei M. Khokhlov
Department of Astronomy
University of Texas at Austin
Austin, TX 78712
phone: 512-471-3397
fax: 512-471-6016
e-mail: a [email protected]
Submitted to Combustion and Flame, February 1995.
Revised version submitted December 1995.
Accepted March 1996.
1
A Theory of DDT in Unconfined Flames
Alexei M. Khokhlov,* Elaine S. Oran,** J. Craig Wheeler*
* Department of Astronomy
University of Texas at Austin
Austin, TX 78712
** Laboratory for Computational Physics
Naval Research Laboratory
Washington, DC 20375
Abstract
This paper outlines a theoretical approach for predicting the onset of detonation in uncon-
fined turbulent flames. Two basic assumuptions are made: 1) the gradient mechanism is
the inherent mechanism that leads to DDT in unconfined conditions, and 2) the sole mech-
anism for preparing the gradient in induction time is by turbulent mixing and local flame
quenching. The criterion for DDT is derived in terms of the one-dimensional detonation
wave thickness, the laminar flame speed, and the laminar flame thickness in the reactive
gas. This approach gives a lower-bound criterion for DDT for conditions where shock
preheating, wall effects, and interactions with obstacles are absent. Regions in parameter
space where unconfined DDT can and cannot occur are determined.
1. Introduction
The quantitative prediction of deflagration-to-detonation transition (DDT) in energetic
gases is one of the ma jor unsolved problems in combustion and detonation theory. Predict-
ing the occurance of DDT has practical importance because of its destructive potential.
It is also an extremely interesting and difficult scientific problem because of the complex
nonlinear interactions among the different contributing physical processes, such as turbu-
lence, shock interactions, and energy release. An early description of experiments on DDT
is given by Brinkley and Lewis [1], who also describe Karlovitz’s theory [2]. Much of this
theory has subsequently been experimentally confirmed and expanded upon by Oppenheim
and coworkers [3–5]. Excellent reviews that summarize our understanding to date have
been given by Lee and Moen [6] and, most recently, by Sheppard and Lee [7]. Other useful
summaries of mechanisms of DDT have been given by Lewis and von Elbe [8] and Kuo [9].
2
Turbulence plays an important role in DDT. Several apparently different mechanisms
for the DDT in confined conditions have been described, each including the turbulence of
the flame and formation of shocks. On large scales, turbulence deforms the flame front
and increases its surface area. On small scales, it broadens the flame front and causes
mixing. The result is an extended turbulent “flame brush” in which a series of explosions
occurs, one of which finally leads to a detonation. Other routes to detonation may include
an explosion in the boundary layer, or an explosion inside the region between the leading
shock and flame brush.
It is believed that, in most cases, the intrinsic mechanism triggering a detonation is
the explosion of a nonuniformly preconditioned region of fuel in which a spatial gradient
of induction time has been created either by turbulent mixing, shock heating, or both.
This gradient mechanism, first suggested by Zeldovich and collaborators for nonuniform
temperature distributions [10,11], was subsequently found in photo-initiation experiments
by Lee at al. [12], who called it the SWACER mechanism. This mechanism has since
been studied and described extensively (see, for example, [6,13–16]). The mechanisms for
preconditioning the region, that is, the mechanism for preparing an explosive mixture that
has a gradient in induction times, may differ in different situations. It may be created by
a shock wave, turbulence, photo-irradiation, intrinsic flame instabilities, rarefaction, or a
combination of several of these.
It appears to be very difficult to obtain DDT in unconfined conditions [17–19]. This can
be attributed to the geometrical effects of expansion: shocks which precede a deflagration
might be weakened, or turbulence might be damped too much by the expansion, and so
become unable to precondition the mixture. Wagner and coworkers [17] report experiments
in which deflagrations were forced to DDT by passing through screens of specified mesh
sizes. The screens created turbulence of the required scale and intensity. These experiments
suggest that an unconfined deflagration could make the transition to detonation under the
right conditions. This possibility has been suggested for very large vapor clouds [6,20].
A related problem that has been studied experimentally is initiation of detonations
by turbulent jets [18,21–24]. In these experiments, a jet of hot product gases in injected
into an unburned, cold mixture. The turbulence generated by the interaction of this jet
and the background gas created a nonuniform, preconditioned region in which detonation
may occur. For these experiments, the effects of reflected shocks and interaction with walls
is minimal compared to DDT in tubes. Therefore, these experiments provide important
information on the critical size of the region capable of triggering DDT.
We then ask the following question: What are the minimal requirements for DDT in an
3
idealized situation when all wall effects and incident shocks are eliminated? If we can answer
this question, we have a lower bound for DDT conditions. Knowing the necessary conditions
for unconfined DDT, we may then draw conclusions about the relative importance of wall
effects and shocks of different strengths. One possible application of this theory could be
to create reproducible detonations in the shortest time and smallest space, as required for
pulse-detonation engines. Another application is to the theory of supernovae. If DDT does
occur in supernovae, as indicated by observations [25,26], it would arise from an unconfined
transition. Currently, there is no quantitative theory explaining exactly how and when an
unconfined transition would occur.
In this paper, we derive a theory for unconfined DDT. That is, we address the situation
where there are no external or reflected shocks, and no wall effects. We make two basic
assumptions:
i. The gradient mechanism is the inherent mechanism that leads to DDT in unconfined
conditions, and
ii. The sole mechanism for preparing the gradient in induction time is by turbulent mixing
and local flame quenching. By this assumption, the role of turbulence is to mix high-
entropy products of burning and low-entropy unreacted fuel. Such mixing creates
gradients of temperature and concentration which have opposite signs. Turbulence-
generated shocks are ignored.
Given these assumptions, there are two fundamental questions to address: 1) What is
the minimum size of a mixed region capable of generating a detonation, and 2) What level
of turbulence is required to create this region? We address these two questions separately,
and then combine the answers to derive the conditions for unconfined DDT. Here we do not
address the question how these conditions may be produced, but give the scale and intensity
of the turbulence that is required. The derived criterion gives lower bounds on conditions
for DDT that does not take into account many secondary effects that may facilitate DDT.
We then conclude with a discussion of the quantitative importance of secondary effects.
4
2. Critical Size of the Preconditioned Region
In this section we address the first of the two questions formulated in the introduction. We
consider the process of the initiation of detonation that arises from the explosion of reactive
gas with a nonuniform distribution of induction times. We assume that the nonuniformity is
a result of mixing of high-entropy products and low-entropy unreacted fuel. We determine
the minimum size Lc of a mixed region capable of triggering a detonation. Whether and
how such a region can be created is a separate question that is studied in Section 3.
We can imagine a variety of regions of different shapes and degrees of mixing created
by turbulence. Here we consider the simplest representative case of a mixed region with a
linear one-dimensional distribution of products. In the future, we plan to consider regions
with different shapes, and thus explore the influence of geometry. However, we do not
believe geometrical considerations will qualitatively change our conclusions (Section 3.2).
2.1 Spontaneous Burning
To facilitate the discussion of a nonuniform explosion of a mixed region, it is useful to
discuss in general terms the idea of spontaneous burning. This concept was first introduced
in Zeldovich and Kompaneetz [27]. Consider a mixture with a nonuniform distribution of
temperature T (x) and chemical composition Y (x). The induction time becomes a function
of spatial coordinate, τ (T (x), Y (x)). In the absence of any physical communication between
different fluid elements, the explosion will start at a point of minimum τ , and then will
spread spontaneously with a “phase” speed
dx (cid:19)−1
Dsp = (cid:18) dτ
which can have any value from zero to infinity. A spontaneous reaction wave does not
(1)
,
require any physical agent in order to spread. Therefore, its speed is not limited by the
speed of light. In reality, there is physical communication between fuel elements. If the
spontaneous velocity is too small, shocks and even heat conduction may cause faster flame
propagation than that prescribed by equation (1).
Let δ t be the time during which the bulk of chemical energy is released after the
induction period is over, δ t << τ . We can define the thickness of the spontaneous wave as
lsp ≃ Dsp δ t. If Dsp → ∞, the thickness of the wave also goes to ∞. This corresponds to a
constant volume explosion. If Dsp is comparable to the speed of a detonation, on the other
hand, its thickness is also comparable to the thickness of a detonation wave. In this latter
case, δsp may become much less than the size of the system under consideration. Then the
5
spontaneous wave may be viewed as discontinuity which obeys the Hugoniot relations for
a discontinuity with energy release.
On a pressure – specific volume plane, spontaneous burning is represented by points
located on a detonation adiabat. This is shown on Figure 1a, where the regimes of spon-
taneous burning occupy the part of the detonation adiabat from point I to point CJ. The
position of the spontaneous regime on the adiabat is determined by the intersection of
the Rayleigh line dP /dV = −(Dsp/V0 )2 with the detonation adiabat. The regime I cor-
responds to an infinite spontaneous velocity when all matter burns simultaneously due to
uniform preconditioning. The point CJ corresponds to the minimum possible velocity of a
steady spontaneous wave, and is equal to the Chapman-Jouguet velocity of a steady deto-
nation. The same part of the detonation adiabat, I – CJ, is occupied by weak detonations.
The difference between spontaneous waves and detonations, is that there is no shock wave
present inside a spontaneous wave. The structure of a Chapman-Jouget detonation and a
spontaneous wave of the same strength is shown schematically in Figure 1b.
In a detonation, the material is first shocked (point S in Figure 1), and then expands
towards the CJ point along the S – CJ line. In the corresponding spontaneous wave, the
material is continuously compressed along the O – SP line until it reaches the CJ point (or
some other point SP). The pressure, density, and velocity in a spontaneous wave become
larger than those of a constant volume explosion (point I) because burning does not proceed
simultaneously! There exists a pressure gradient inside the wave pointing opposite to its
direction of propagation, since at any instant the wave consists of fluid elements with
different amounts of released energy. As a result, a fluid element passing through the wave
is compressed and accelerated by this gradient. The slower the wave moves, the longer
is the time spent inside the wave, and the greater are the pressure, density and velocity
jumps across the wave. The principle of causality is not violated in the spontaneous wave,
as explained in [28]. Although the speed of the spontaneous wave is a phase speed, it is a
real supersonic wave of burning which looks like a detonation in terms of the hydrodynamic
parameters of burned material.
We have discussed the situation where the spontaneous wave speed is greater or equal
to the CJ detonation velocity, DCJ . Suppose the gradient in induction time is such that
Dsp is initially greater than DCJ , but then it decreases so that it becomes less than DCJ .
In this case, when the spontaneous wave crosses the CJ threshold, the burned material
immediately behind the wave, which moves with the local sound speed relative to the wave,
will tend to overcome the wave and produce a shock. Consider an intermediate regime with
such a shock, O – O’ – S’ – CJ, shown in Figure 1. First, material burns in a spontaneous
6
wave from O – O’, then it is shocked to point S’, and then returns to the CJ point. The
transition from the spontaneous wave to a CJ detonation may then proceed through a
sequence of such regimes, with increasing shock strengths.
The description given in the last paragraph is a quasi-steady picture that is applicable
only if the spontaneous wave velocity changes slowly enough.
If the spontaneous wave
velocity changes too fast, that is, the gradient is too steep, the shock and reaction will
separate, and the CJ detonation will not form. In the process of transition from spontaneous
wave to CJ detonation, the spontaneous velocity must change slowly enough so that the
shock and reaction do not separate. This means that the nonuniform region must be
large enough so that this separation does not occur, and this, in turn, gives a criterion for
unconfined DDT.
2.2 Formulation of the Problem
Consider an idealized one-dimensional system with the equation of state P = (γ − 1)Et and
T = P /ρ, where P , T , ρ and Et are the pressure, temperature, mass density and thermal
energy density, respectively. The chemical reaction is described by a first-order Arrhenius
expression,
= −Y exp (cid:18)−
T (cid:19) ,
Q
where Q is the activation energy, and the chemical variable Y ranges from Y = 1 for pure
d Y
dt
(2)
reactants to Y = 0 for total products. Units of distance and time are such that the pre-
exponential factor in equation (1) is unity, and the gas constant is R = 1. Planar geometry
is assumed. The system obeys the Euler equations,
+
∂ρ
∂ t
∂ρU
∂ t
∂E
∂ t
(ρU ) = 0 ,
∂
∂x
∂
∂x (cid:0)ρU 2 + P (cid:1) = 0 ,
∂
( (E + P ) U ) = 0 ,
∂x
+
+
(3)
where U is the fluid velocity, E = Et + ρU 2 /2 − ρqY is the total energy density including
chemical energy, and q is the total energy release per unit mass.
The initial temperature and density of the fuel are T0 and ρ0 . The products of isobaric
burning, an approximation to burning in a laminar flame, have a temperature T1 = T0 +
γ (cid:17). By our assumption, we consider a nonuniform region created by mixing the
q (cid:16) γ−1
products of isobaric burning and fresh fuel, such that there is a linear spatial distribution
7
of reactants Y (x) and temperature T (x),
Y (x) = (cid:26) x/L,
0 ≤ x ≤ L
1,
x > L
T (x) = T1 − (T1 − T0 ) Y (x) ,
where L is the size of the mixed region. Initially, the velocity of the material is zero, and
(4)
the pressure P0 is constant everywhere. The boundary conditions at x = 0 are reflecting
walls (symmetry conditions).
The system is prepared in an initial state and then evolves in time, first until ignition
takes place, then to the formation (or failure) of detonation, and then to the time when
the generated detonation or shock leaves the computational domain. The cases considered
are listed in Table 1. Parameters for the standard case H1 with P0 = 1 and T0 = 1 are
chosen to approximate a detonation in a stoichiometric hydrogen and oxygen mixture at
pressure of 1 atm and temperature of 293 K [29,30]. The extra cases, H2 (T0 = 2) and H3
(T0 = 3), are considered to study the sensitivity of the detonation formation to the initial
temperature of the fuel.
The system of equations (2) and (3) is integrated numerically using a one-dimensional
version of a time-dependent, compressible fluid code based on the Piecewise Parabolic
Method (PPM) [31,32]. PPM is a second order Godunov-like method which incorporates
a Riemann solver to describe shock waves accurately. Shocks are typically spread on one
or two computational cells wide. A piecewise parabolic advection algorithm advects sharp
shockless features, such as density and composition discontinuities or gradients, without
diffusing them excessively or changing their shape. Contact discontinuities are typically
kept two or three cells wide. Details of the implementation are given in [33,34]. The chem-
ical reaction is coupled to fluid dynamics by time-step splitting. The kinetic equation (2)
is integrated together with the equation of energy conservation using adjustable substeps
to keep the accuracy better than 1%. The grid spacing is selected so that there are at least
10 cells within a detonation wave reaction zone. The convergence of numerical solutions
was thoroughly tested by varying the number of computational cells from 1024 to as many
as 65536 in some cases.
2.3 Detonation Formation Inside the Mixed Region
as
(5)
The induction time τ as a function of temperature T and fuel fraction Y can be expressed
(γ − 1)qY (cid:18) T 2
Q (cid:19) exp (cid:18) Q
T (cid:19)
1
8
τ (T , Y ) ≃
using the Frank-Kamenetskii approximation [35], valid when T /Q << 1, and assuming the
induction takes place at constant volume. The derivatives of τ with respect to T and Y are
∂ τ
∂T ≃ −
τ Q
T 2 ,
∂ τ
∂Y ≃ −
τ
Y
.
(6)
For the mixture considered here, the values of T and Y are related by equation (4). The
function τ (T , Y (T )) then has a minimum at Tm , found by solving dτ /dT = ∂ τ /∂T +
(∂ τ /∂Y )(d Y /dT ) = 0, so that
T 2
m + QTm − QT1 = 0 .
(7)
This gives
(8)
2Q (cid:19) ,
T1
Tm = Q q1 + 4 T1
Q − 1
≃ T1 (cid:18)1 −
2
xm = L (cid:18) T1 − Tm
T1 − T0 (cid:19) ≃
LT1
,
2Q
for T0 << T1 . The point xm is the first to ignite. From this point, a spontaneous reaction
wave propagates with the speed
dx (cid:19)−1
T 2 (T − T1 )
Dsp = (cid:18) dτ
(T1 − T0 )(T 2 + QT − QT1 )
By virtue of equation (7), the speed of the reaction wave is infinite at point xm . Thus,
the reaction wave initially propagates supersonically, as described in Section 2.1 We are
interested in the propagation of the wave to the right, x → L, where the energy released
by the wave increases. The velocity of the wave decreases towards larger x, and becomes
=
L
τ
.
(9)
equal to the local sound speed at some point xs determined by
Dsp (xs ) = DCJ .
(10)
At this point, a pressure wave forms which runs into the mixture ahead of the decelerating
reaction wave. Whether this pressure wave is strong enough to accelerate burning and to
evolve into a detonation wave depends on the length L of the mixed region.
There are two processes involved in the transformation of the pressure wave into a
detonation. First, the pressure wave must steepen into a shock. This shock must accelerate
burning so that a shock–reaction complex forms. Second, the shock–reaction complex must
survive the propagation down the temperature gradient. We denote as Ls the first critical
length of the mixed region such that for L < Ls the shock–reaction complex does not form.
9
For L > Ls , the shock–reaction complex successfully forms within the mixed region. We
denote as Ld the second critical length of the mixed region such that for L > Ld , the shock–
reaction complex survives and passes as a detonation into the cold fuel. For Ls < L < Ld ,
the shock–reaction complex dies inside the mixed region.
Values of Ls and Ld were determined by the numerical simulations described in Section
2.2 by performing a series of simulations in which the size L of the mixed region was
varied. Figures 2–5 show results of simulations for H1 for four different choices of L. Each
figure shows the evolution with time of the pressure, velocity, temperature, and reactant
concentration. Figure 2 shows the results for the smallest mixed region, L = 30xd , where xd
is the half-reaction width of a CJ detonation. This region is so small that the quasi-steady
spontaneous wave cannot form. The pressure wave is too weak to form a shock–reaction
complex. The pressure wave generated by the spontaneous burning steepens into a shock
outside of the mixed region.
For L = 300xd , shown in Figure 3, a shock wave forms at the point predicted by
equation (10), and the complex forms. The simulations show that the shock-reaction com-
plex is far from a steady CJ detonation and cannot be described as a small quasi-steady
deviation from the CJ state. The peak pressure is at least a factor of two less than the von
Neumann pressure for the equivalent CJ detonation at the local condition. Figure 3 shows
that, soon after the complex is formed, the reaction zone and shock wave separate, and
only a shock wave leaves the mixed region. This is because the shock–reaction complex,
after being formed, must propagate through the mixture with continuously decreasing tem-
perature. The temperature gradient causes rapid decrease in the postshock temperature,
and, consequently, rapid growth of the induction time in the postshock material.
Figures 4, L = 500xd < Ld , shows a case similar to Figure 3, but the shock–reaction
complex decouples close to the end of the mixed region. In Figure 5, L = 960xd > Ld ,
the complex transforms into a detonation, and passes into the cold unmixed fuel. The
critical condition for the initiation of detonation in mixed fuel and products is that the
shock–reaction zone complex survives its propagation through the temperature gradient.
The critical lengths, Ld , of a region capable of triggering a detonation, as determined by
such simulations, are presented in Table 1 for cases H1 – H3.
The value of the critical length Ld is sensitive to initial temperature T0 . An increase
of T0 facilitates the initiation of detonation. Cases H2 and H3 in Table 1 show that Ld
decreases by a factor of six if the initial temperature is tripled. This can be explained if
the criterion for the detonation formation is not the creation, but rather the survival of the
shock–reaction complex. For higher initial temperature, the postshock induction time is
10
less sensitive to variations of background conditions (see equation 6), and so it is easier for
the shock–reaction complex to adjust to changing conditions.
11
2.4 Relation to Jet Initiation Experiments
One possible check on the theory described above for determining Ld is to compare the
predictions of Section 2.3 with the results of turbulent jet-initiation experiments [18,21–24].
In these experiments, a jet of hot product gases in injected into an unburned, cold mixture.
The jet can be characterized by the size of the orifice, d, through which hot products are
injected. The turbulence caused by the interaction of this jet and the background gas
creates a nonuniform, preconditioned region in which detonation may occur. The largest
scale of the turbulence and the size of the mixed region are also characterized by d. For
these experiments, the effects of reflected shocks and interaction with walls is believed to
be small. The velocity of the jet is approximately sonic with respect to the unburned
background material, Thus the strengths of the shocks formed by the exiting jet resulted in
overpressures in the unburned gas of about a factor of two or less. The temperature increase
was small. Ignition occurred in the jet and seemed to be unaffected by wall interactions.
Depending on d, two distinct ignition regimes were found. For small d, deflagrations
were ignited at many points inside the mixed region. There was no transition to detonation.
For larger d, there was an explosion in the mixed region that led to detonation. From these
experiments, the minimum value of d for which DDT occurred was d > 10 − 20lc , where lc
is the detonation cell size.
The half-reaction zone length xd , in terms of which we derived our estimates of Ld ,
is a theoretical parameter. What is measured in experiments is a detonation cell size lc .
In order to estimate lc for the case H1, we use the results of two-dimensional simulations
of detonation cell structure for conditions similar to H1 [30]. Scaling the results of these
simulations to nondimensional units, we find lc ≃ 27xd , where we have taken lc to be
the height of a detonation cell. That is, the critical size of the mixed region in case H1 is
Ld ≃ 36lc . Thus the theoretical estimate of Ld is in qualitative agreement with experiments.
The somewhat larger theoretical value, 36lc compared to 10 − 20lc , could be the result of
the simplifying assumptions (one-step kinetics, neglect of multi-dimensional effects) made
in this paper.
There have been other efforts to relate Ld to lc . For example, Knystautus et al.
[18] found that Ld ≃ 13lc based on the analogy between DDT and direction initiation
of detonation by an energy source. Dorofeev et al. [21] report Ld > 7lc based on their
computations.
12
3. Critical Turbulent Velocity for DDT
3.1 Preconditioning by Turbulence
The discussion in the previous section established that the size of the region required
to trigger a detonation is large compared to the one-dimensional detonation thickness,
Ld ≃ 103xd for case H1. Now the question is how to create this region. In an unconfined
space, turbulence is the only mechanism available. The turbulence in the region of size
Ld must be strong enough to create microscopic mixing in this region. Turbulence on
large scales must be intense enough to pack individual laminar flame sheets close together.
Turbulence on small scales must be strong enough to broaden and destroy individual flame
sheets so that the products and fuel can mix to form a microscopic region with a gradient
of induction times.
There are generally two regimes of turbulent flames we need to consider. The first is a
regime of multiple flame sheets, in which the flame is irregular on large scales but laminar
on small scales. The second is the distributed burning regime, in which the turbulence
is so strong that it modifies the laminar flame structure (See, for example [36,37]. The
transition between the multiple flame sheet and distributed burning regimes represents the
condition where the creation of the large-scale nonuniform distribution of induction times
becomes possible. The flame will be affected by the turbulence on scales λ ≥ λG , on which
the turbulent velocity is greater than or equal to the laminar flame speed, Sl . Here λG is
the Gibson scale defined by the condition
U (λG ) = Sl
(11)
where U (λ) is the turbulent velocity on the scale λ. The transition between the two
turbulent regimes happens approximately when λG approaches the thickness of the laminar
flame xl [36]. This estimate is approximate and does not account for the effects of viscosity,
which becomes important when λG approaches the viscous microscale λK . The viscosity
destroys turbulent eddies of size xl . Poinsot et al. [38] have shown theoretically that because
of this effect, eddies larger than λG with velocitiy greater than Sl are needed to quench the
flame. This has been substantiated by the experimental work by Roberts and Driscoll [39]
who showed that eddies a factor of four larger are required.
Consider, for simplicity, a Kolmogorov cascade inside the turbulent flame brush such
that on the scale λ, the turbulent velocity is
L (cid:19)1/3
Uλ ≃ UL (cid:18) λ
,
13
(12)
where L is the driving scale of the turbulence, which could be approximately equal to or
larger than the size of the turbulent flame brush, and UL is the turbulent velocity on this
scale. In this case, the Gibson scale λG becomes
UL (cid:19)3
λG ≃ (cid:18) Sl
L .
The condition λG = xl now can be used to define the intensity of the turbulent motions
needed for DDT,
(13)
xl (cid:19)1/3
UL = K Sl (cid:18) L
where we introduced a coefficient K ≃ 1 which describes the ability of the laminar flame
to survive stretching and folding caused by turbulence on scales of the order of xl . Once
the condition of equation (14) is reached for L ≥ Ld , DDT can occur by the gradient
mechanism.
(14)
,
For a typical flame, the thickness of the laminar flame xl is approximately an order of
magnitude less than xd . That is, Ld ≃ 104xl . For a flame with Ld /xl = 104 , the intensity
of turbulent motions required for DDT on the scale of Ld must be about ULd ≃ 20Sl , as
follows from equation (14). For example, consider an equimolar acetylene-oxygen flame
with a laminar flame speed of 5 m/s [40]. ¿From equation (12), the critical intensity of
turbulent motions is approximately UL ≃ 100 m/s, The critical turbulent velocity could be
considerably less in confined conditions because of the presence of shocks.
In unconfined situations, there are two possible sources of turbulence, the Landau-
Darrieus (L-D) instability and the Rayleigh-Taylor (R-T) instability. The L-D instability
is an intrinsic hydrodynamic flame instability that does not require external acceleration.
The intensity of the L-D induced turbulent motions is unlikely to be much larger than
Sl because of nonlinear stabilization effects [41]. The L-D instability is thus not likely
to produce the level of turbulence required for DDT in any reasonable conditions. The
characteristic turbulent velocity associated with the R-T instability on scale L is of the
order of ≃ √gL ≃ 3√L m/s for L in meters. The level of turbulence required for DDT can
thus be achieved only on scales of ∼ 100 m. This could explain why DDT in unconfined
situations is so rarely observed. To obtain DDT in the laboratory, we need some other way
of inducing much higher turbulent intensities.
3.2 Secondary Effects
When a region smaller than Ld ignites, it can still generate a substantial shock. The
dependence of the maximum shock pressure on L found from the simulations is shown in
14
Figure 6. The shock strength is high for L larger than, say, 0.5Ld , but rapidly decreases
for smaller L. There are two possible effects these shocks may produce, one related to the
temperature increase and another to vorticity created by the shocks.
The shock may raise the temperature in a region of the mixture that is about to
explode, and this may facilitate the survival of the shock–reaction complex. Table 1 shows
that the increase of the initial temperature from T0 = 1 to T0 = 2 decreases Ld by a factor
of four. The increase of the initial temperature by a factor of two requires, however, a shock
strength Ps/P0 ≃ 8. This shock strength can be provided only by explosions of regions of
size L > 0.5Ld (see Figure 6). That is, this effect may slightly decrease the one-dimensional
estimate of Ld , but is not likely to change it drastically.
Another effect of a failed initiation on the surrounding material might be the baroclinic
generation of additional vorticity [1,42]. Such a secondary source of turbulence reduces the
amount of turbulence that must be generated by the primary sources. The turbulent
velocity induced by this mechanism may be of order of the postshock velocity. This source
of secondary vorticity may be very important in facilitating DDT, but only when the
conditions are already close to critical. The amount of secondary vorticity will rapidly
decrease with decreasing L. We conclude that our estimate of Ld may decrease by a factor
of about two, but will not change drastically if the baroclinic mechanism is taken into
account.
The ma jor uncertainty in the estimation of the required turbulent velocity comes from
our lack of exact knowledge of flame behavior on scales ∼ xl in the turbulent velocity field.
The standard definition of the Gibson scale as the scale at which the turbulent velocity is
equal to the laminar flame speed, UλG = Sl , and the assumption that microscopic mixing
begins when λG = xl , gives K = 1 in equation (14). As mentioned above, recent work by
Poinsot et al. [38] and Roberts and Driscoll [39] suggest K > 1. There is also some evidence
from numerical simulations of turbulent flames that this coefficient might be K ≃ 3 − 5,
which would increase the critical turbulent velocity accordingly [33]. This must be studied
in future numerical simulations and experiments.
The same kind of mixing and flame quenching must also take place in the flame brush
of a turbulent deflagration in a tube in order to have DDT in a confined situation. Although
shock preconditioning definitely plays an important role in confined situations, there should
be a qualitative similarity between triggering detonation by the explosion in the middle of
the brush and DDT in unconfined conditions. Carefully planned experiments on DDT
in tubes with quantitative characterization of the the turbulent velocity field prior to the
explosion in the brush might be used to shed light on the exact value of coefficient K .
15
4. Conclusions
There are two key elements to the theory presented above for unconfined DDT:
1. The size of the region Ld that can trigger DDT in a mixture of hot burning product and
fuel. We estimate that Ld ∼ 103xd , where xd is the thickness of the one-dimensional
reaction zone of the Chapman-Jouguet detonation, or Ld ≃ 36 lc , where lc is the
detonation cell size, or Ld ≃ 104xl , where xl is the laminar flame thickness. This
implies that large-scale mixing is required to precondition the region.
2. The intensity of turbulent motions required for the region of size Ld to undergo DDT.
This is estimated from the requirements that the Gibson scale inside this region be
comparable to or less than the thickness of the laminar flame (equation 14). This
requires the speed of the turbulent flame brush to be ∼ 102 times faster than the
laminar flame.
The criterion of DDT in unconfined flames given here can be formulated in terms of the
following three parameters of a reactive gas: the one-dimensional thickness of a CJ detona-
tion, xd , the velocity Sl , and the thickness xl of the laminar flame. The critical size of the
mixed region Ld can be directly related to xd (Section 2.3), and the latter two parameters
determine the critical intensity of turbulence in the mixed region required for triggering
DDT (Section 3.1).
The high turbulent velocity required for unconfined DDT is extremely difficult to
achieve by turbulence generated by the flame itself or by the Rayleigh-Taylor instability,
which explains why DDT in unconfined flames is so hard to observe. The critical size of
the region Ld derived in this paper is in agreement with the results of hot jet initiation
experiments. The theory may also be extended to confined DDT in the cases when the
explosion leading to detonation takes place in the middle of a turbulent flame brush.
Acknowledgments
This work was sponsored by the National Science Foundation, Texas Advanced Research
Pro ject, the Naval Research Laboratory through the Office of Naval Research, and the
NASA Astrophysical Theory Program. The authors are grateful to C.J. Sung and C.K. Law
for the useful information on laminar flame speeds, to Vadim Gamezo for his unpublished
results of detonation cell simulations in hydrogen-oxygen mixtures, to Martin Sichel for
the discussion of general properties of DDT, and to James F. Driscoll for references and
discussion of the cutoff scales. The computations were performed at the High-Performance
Computing Facility of the University of Texas.
16
Table 1
Table – Simulated Cases
Case T0
γ
H1
H2
H3
1
2
3
1.333
1.333
1.333
q
24
24
24
Q
28.3
28.3
28.3
T1
7.0
8.0
9.0
Ts
5.8
7.1
8.3
Td
10.3
11.4
12.5
xd
51.2
32.3
23.5
Ls/xd
∼ 2 × 102
-
-
Ld /xd
9.5 × 102
3.3 × 102
∼ 2 × 102
T0
γ
q
Q
T1
Ts
Td
xd
Ls
Ld
Initial fuel temperature.
Adiabatic index.
Total chemical energy release.
Activation energy.
Temperature of products of isobaric burning.
Postshock temperature in a Chapman-Jouguet detonation.
Temperature of Chapman-Jouguet detonation products.
Half reaction zone length of Chapman-Jouguet detonation.
Critical length for shock–burning synchronization.
Critical length for detonation survival in cold fuel.
17
References
1. Brinkley, S.R, Jr. and Lewis, B., Seventh Symposium (International) on Combustion,
The Combustion Institute, Pittsburgh, pp. 807–811, 1959.
2. Karlovitz, B.,Selected Combustion Problems, p. 176, London, Butterworths, 1954.
3. Urtiew, P. and Oppenheim, A.K., Proc. Roy. Soc. Lond. A, 295:13–28 (1966).
4. Laderman, A.J., Urtiew, P.A. and Oppenheim, A.K., Ninth Symposium (International)
on Combustion, The Combustion Institute, Pittsburgh, p. 265, 1963.
5. Oppenheim, A.K., Laderman, A.J. and Urtiew, P.A., Combust. Flame, 6:193–197 (1962).
6. Lee, J.H.S. and Moen, I.O., Prog. Energy Combust. Sci., 6:359–389 (1980).
7. Shepherd, J.E. and Lee, J.H.S., Ma jor Research Topics in Combustion, pp. 439–490,
eds. M.Y. Hussaini, A. Kumar, and R.G. Voigt, Springer, New York, 1992.
8. Lewis, B. and von Elbe, G., Combustion, Flames, and Explosions of Gases, Academic,
1987. 3rd edition, pp. 566-573.
9. Kuo, K.K., Principles of Combustion, Wiley, New York, 1986.
10. Zel’dovich, Ya.B., Librovich, V.B., Makhviladze, G.M. and Sivashinsky, G.I., in 2nd
International Colloquium on Explosion and Reacting Systems Gasdynamics, 1969, Aug.
24–29. Novosibirsk, p.10.
11. Zel’dovich, Ya.B., Librovich, V.B., Makhviladze, G.M. and Sivashinsky, G.I., Astro-
naut. Acta 15:313–321 (1970).
12. Lee, J.H.S., Knystautas, R. and Yoshikawa, N., Acta Astronaut. 5:971–982 (1978).
13. Yoshikawa, N., Coherent Shock Wave Amplification in Photochemical Initiation of
Gaseous Detonation, Ph.D. Thesis, Department of Mechanical Engineering, McGill Uni-
versity, 1980.
14. Zel’dovich, Ya.B., Gelfand, B.E., Tsyganov, S.A., Frolov, S.M. and Polenov, A.N.,
Progr. Astronaut. Aeronaut. 114:99-123 (1988).
15. He, L. and Clavin, P., Twenty-Fourth Symposium (International) on Combustion, The
Combustion Institute, Pittsburgh, pp. 1861–1867, 1992.
16. Weber, H.-J., Mack, A. and Roth, P., Combust. Flame 97:281–295 (1994).
17. Wagner, H.Gg., in Proc. Int. Specialist Conf. Fuel-Air Explosions, U. Waterloo Press,
77–99 (1981).
18
18. Knystautas, R., Lee, J.H.S., Moen, I.O. and Wagner, H.Gg., Proceedings of the 17th In-
ternational Symposium on Combustion, The Combustion Institute, Pittsburgh, pp. 1235–
1245, 1978.
19. Boni, A.A, Chapman, M., Cook, J.L. and Schneyer, J.P., in Turbulent Combustion,
ed. L. Kennedy, pp. 379–405, (1978).
20. Moen, I.O., Lee, J.H.S., Hjertager, B.H., Fuhre, K. and Eckhoff, R.K., Combust. Flame
47:31–52, (1982).
21. Dorofeev., S.B., Bezmelnitsin, A.V., Sidorov, V.P., Yankin, J.G., and Matsukov, I.D.,
Fourteenth International Colloquium on Dynamics of Explosions and Reactive Systems,
University of Coimbra, Coimbra, 1993, Vol. 2, pp. D2.4.1–D2.4.10.
22. Carnasciali, F., Lee, J.H.S., and Knystautas, R., Combust. Flame 84: 170–180 (1991).
23. Moen, I.O., Bjerketvedt, D., and Jenssen, A., Combust. Flame 61: 285–291 (1985).
24. Medvedev, S.P., Polenov, A.N., Khomik, S.V., and Gelfand, B.F., Twenty-Fifth Sym-
posium (Internqational) on Combustion, The Combustion Institute, Pittsburgh, pp. 73–78,
1994.
25. Khokhlov, A.M., Astronomy ans Astrophys. 245:114 and 245:L25 (1991).
26. Hoflich, P.A., Khokhlov, A.M. and Wheeler, J.C. Astrophys.J. 444:831–847 (1994).
27. Zeldovich, Ya.B., and Kompaneets, S.A., Theory of Detonations, New York, Academic
Press, 1960.
28. Khokhlov, A.M., Astron. Astrophys. 246: 383–396 (1991).
29. Strehlow, R., Maurer, R.E. and Ra jan, S., AIAA 7:323–328 (1969).
30. V. Gamezo (private communication).
31. Colella, P., and Woodward, P.R., J. Comp. Phys. 54:174–201 (1984).
32. Colella, P., and Glaz, H.M., J. Comp. Phys. 59:264–289 (1985).
33. Khokhlov, A.M., Astrophys.J. 449:695–713 (1995).
34. Khokhlov, A.M., Oran, E.S. and Wheeler, J.C., Combust. Flame (1996) 105:28–34.
35. Frank-Kamenetskii, D.A., Diffusion and Heat Transfer in Chemical Kinetics, Nauka,
Moscow (1967).
36. Peters, N., in Numerical Approaches to Combustion Modeling, eds. E.S. Oran and J.P.
Boris, AIAA, Washington, DC 1991, pp. 155-182.
37. Williams, F.A., in Mathematics of Combustion, ed. J.D. Buckmaster, Vol. 2, of Frontiers
in Applied Mathematics, SIAM, Philadelphia, pp. 97–131 (1985).
19
38. Poinsot, T., Veynante, D., and Candel, S., J. Fluid Mech. 228: 561–606 (1991).
39. Roberts, W.L., and Driscoll, J.F., Combust. Flame 87: 245–256 (1991).
40. Sung, C.J. (private communication).
41. Zel’dovich, Ya.B., Journ. Applied Mech. Tech. Phys. 7:68 (1966).
42. Picone, J.M., Oran, E.S., Boris, J.P. and Young, T.R. Jr., in Dynamics of Shock
Waves, Explosions, and Detonations, eds. J.R. Bowen, N. Manson, A.K. Oppenheim, and
R.I. Soloukhin, 94, Prog. Astronaut. Aeronaut., 1985, pp. 429–488.
20
Figure captions
Fig. 1 – (a) Regimes of burning in the P − V plane, (V = 1/ρ). Solid curve: detonation
adiabat. Dashed curves: shock adabats. O , initial state of cold fuel; I , state of products of
constant volume explosion; C J , Chapman-Jouguet state; S , unburned post-shock state for
CJ detonation; S ′ , post-shock state for intermediate regime; O ′ , preshock state for inter-
mediate regime; SP , state of products of spontaneous burning. (b) Schematic of pressure
profile of a CJ detonation, spontaneous wave, and intermediate regime. For the sponta-
neous wave, the pressure changes smoothly from O to C J , in contrast to the discontinuity
O – S in the detonation. In the intermediate regime, pressure increases smoothly from O
to O ′ , and then passes through the shock O ′ – S ′ .
Fig. 2 – Pressure, velocity, temperature, and reactant concentration as a function of dis-
tance at different times (marked by numbers in increasing order) for case H1 (Table 1) with
L/xd = 30. Vertical arrow indicates where Dsp = DCJ . Values of xd are given in Table 1.
Fig. 3 – Same as Fig. 1, but with L/xd = 300.
Fig. 4 – Same as Fig. 1, but with L/xd = 500.
Fig. 5 – Same as Fig. 1, but with L/xd = 960.
Fig. 6 – Maximum shock pressure Ps generated during a nonuniform explosion as a function
of the length L of a mixed region.
21
|
astro-ph/0307352 | 2 | 0307 | 2003-11-18T17:52:51 | Pulsar Recoil by Large-Scale Anisotropies in Supernova Explosions | [
"astro-ph"
] | Assuming that the neutrino luminosity from the neutron star core is sufficiently high to drive supernova explosions by the neutrino-heating mechanism, we show that low-mode (l = 1, 2) convection can develop from random seed perturbations behind the shock. A slow onset of the explosion is crucial, requiring the core luminosity to vary slowly with time, in contrast to the burst-like exponential decay assumed in previous work. Gravitational and hydrodynamic forces by the globally asymmetric supernova ejecta were found to accelerate the remnant neutron star on a timescale of more than a second to velocities above 500 km/s, in agreement with observed pulsar proper motions. | astro-ph | astro-ph |
Pulsar recoil by large-scale anisotropies in supernova explosions
L. Scheck,1 T. Plewa,2, 3 H.-Th. Janka,1 K. Kifonidis,1 and E. Muller1
1Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Str. 1, D-85741 Garching, Germany
2Center for Astrophysical Thermonuclear Flashes,
University of Chicago, 5640 S. Ellis Avenue, Chicago, IL 60637, USA
3Nicolaus Copernicus Astronomical Center, Bartycka 18, 00716 Warsaw, Poland
(Dated: December 6, 2018)
Assuming that the neutrino luminosity from the neutron star core is sufficiently high to drive su-
pernova explosions by the neutrino-heating mechanism, we show that low-mode (l = 1, 2) convection
can develop from random seed perturbations behind the shock. A slow onset of the explosion is cru-
cial, requiring the core luminosity to vary slowly with time, in contrast to the burst-like exponential
decay assumed in previous work. Gravitational and hydrodynamic forces by the globally asymmetric
supernova ejecta were found to accelerate the remnant neutron star on a timescale of more than a
second to velocities above 500 km s−1, in agreement with observed pulsar proper motions.
PACS numbers: 97.60.Bw, 97.60.Gb, 95.30.Jx, 95.30.Lz
Young pulsars are observed to have average space ve-
locities of 200 -- 500 km s−1 with highest values above
1000 km s−1 and still ambiguous hints for a double peak
distribution [1]. There is no clear statistical correlation
with the magnetic moment or rotation of the pulsar, al-
though in two cases (Vela and Crab) the direction of
motion appears to be aligned with the spin axis.
A connection of the pulsar motions with the super-
nova (SN) explosion is suggested by neutron star (NS) --
SN remnant associations and by the properties of binary
systems with one or both components being a NS. Natal
kicks are required, e.g., by the spin-orbit misalignment
and high orbital eccentricities observed in some binaries
(for a review, see [2]).
Various mechanisms have been proposed to explain
these kicks. One possibility invokes asymmetric mass
ejection during the SN [3]. This may be caused by large-
scale density inhomogeneities in the pre-collapse core of
the progenitor star [4] or convective instabilities in the
neutrino-heated layer behind the SN shock [5, 6]. The
kicks could also be a consequence of unequal momentum
fluxes in a jet and an anti-jet that might be linked to the
start of the SN explosion [7] or to the NS formation [2].
Alternatively or in addition, anisotropic neutrino (ν)
emission from the nascent ("proto-") neutron star (PNS)
might transfer the momentum [8]. A "neutrino rocket
engine" of the latter kind could result from the magnetic
field-strength dependence of ν-matter interactions if ex-
tremely strong fields with hemispheric asymmetries build
up in a PNS [9]. The ν transport also depends on the
field direction, e.g., through parity violating corrections
of weak interaction cross sections [10] or due to resonant
flavor transitions [11]. In case of a significant dipole com-
ponent such effects can lead to kicks of a few 100 km s−1
for magnetic fields in excess of 1015 G [2].
In this Letter we present new two-dimensional (2D)
calculations of hydrodynamic instabilities during the on-
set of SN explosions which show that global asymmetries
and the PNS recoil can naturally grow to a sufficient
size without invoking artificial initial conditions, extreme
physical assumptions, or exotic ν physics. Our computa-
tions improve previous ones [6] with respect to numerical
resolution, a full 180o lateral grid, and the treatment of
ν transport, extending them also in the computed evolu-
tion time and model set.
Modeling concepts. We assume that the explosion is
powered by ν-energy deposition between the PNS and
the SN shock [12]. Although the currently most elaborate
numerical models are not able to confirm the viability of
this ν-heating mechanism, it is still the best-studied and
most promising way to explode massive stars [13].
SN theory is currently hampered by our incomplete
knowledge of the nuclear physics and ν interactions in the
dense matter inside the PNS. This implies uncertainties
for self-consistent models of the full problem, e.g. with
respect to the magnitude of the ν fluxes emitted by the
cooling PNS. Therefore, we replace the shrinking, high-
density core of the PNS by a gravitating sphere whose
radius coincides with the contracting, impenetrable inner
boundary of our computational grid.
At this boundary, number and energy fluxes of ν and
¯ν of all three lepton flavors are imposed with chosen ini-
tial values and time dependence. In all simulations these
boundary values were taken to be isotropic and were kept
constant for a chosen period of time after shock forma-
tion (either 0.7 s or 1 s, within which 1/3 of the gravi-
tational binding energy of the nascent NS was assumed
to be radiated away in neutrinos). The grid boundary is
located somewhat below the neutrinosphere of νe. It has
an optical depth which, for typical ν energies, rises from
few initially to several 100 at the end of the simulations.
The use of this inner boundary allows us to explore the
response of the collapsing SN core to different ν lumi-
nosities. Higher values of the latter lead to larger energy
deposition behind the SN shock and therefore to more
powerful explosions. The dominant heating reactions are
2
∂tL + c ∂rL = 4πr2c Q−
ν − κcL . Here L = L(r, t) is
the total ν number flux or ν luminosity, respectively, Q−
ν
the rate of ν loss by the stellar medium per unit vol-
ume, κ ≡ 4πr2c Q+
ν /(Lc) the corresponding absorptiv-
ity, c the speed of light, and c the "effective speed" of ν
propagation, which is governed by diffusion at high den-
sities and reaches c at large radii. The integration for L
in radius r and time t (lateral flux components are ig-
nored) can be done analytically when Q−
ν , κ and c are
assumed to be constant within the cells of the numer-
ical grid (consistent with this, the above equation was
derived by employing ∂tc = 0). Instead of determining
c from the solution of the Boltzmann equation, we use
an analytic representation in terms of the optical depth
that was obtained from fitting results of detailed ν trans-
port in the outer layers of the SN core. Neutrino-matter
interactions via charged-current processes with nucleons,
thermal pair creation, and scattering off nuclei, n, p, e−,
and e+ are evaluated by adopting Fermi-Dirac ν spectra
with a temperature determined by the ratio between ν
energy and number density. The chemical potentials of
the spectra are taken to be equal to the equilibrium val-
ues at high optical depths and approach values near zero
outside of the neutrinospheres.
This approximate ν transport retains the hyperbolic
character of the transport problem, conserves lepton
number and energy globally, and reproduces basic prop-
erties of accurate Boltzmann transport calculations de-
spite of radical simplifications. It is coupled to our 2D hy-
drodynamics code by operator-splitting with a predictor-
corrector step.
Our hydrodynamics code and equation of state are de-
scribed in Ref. [14]. We typically use 400 geometrically
spaced radial zones and one degree lateral resolution.
The ν transport is solved in each angular bin separately.
The 2D simulations were started some milliseconds after
SN shock formation from detailed core-collapse models
with fluid velocities randomly perturbed with an ampli-
tude of typically 0.1%.
Results. Figure 1 shows four sequences of runs that
followed the post-bounce evolution of different 15 M⊙
progenitors for one second. The crosses mark results
based on the use of Model WPE15 [15], the solid dots of
Model LSC15 [16], the open circles of Model s15s7b2 [17],
and the triangles of a model with a structure like the lat-
ter but including rotation [13]. Starting with a rotation
period of 12 s in the iron core [18], the post-collapse core
spins differentially within 10 -- 20 ms and speeds up as it
contracts. The rotation period in the ν-heating layer
varies between some 10 ms and several 100 ms [13].
Model LSC15, in particular, differs from the others by
having significantly higher densities at the edge of the
iron core and in the silicon shell (at a time when the
cores have evolved to the same central density). This
delays the start time, texp, of the explosion, reduces the
explosion energy, Eexp, and leads to a larger NS baryonic
FIG. 1: Explosion timescale texp and, at one second after SN
shock formation, explosion energy Eexp, NS baryonic mass
Mns, gas anisotropy parameter αgas, NS recoil velocity vns,
and NS acceleration ans (from bottom to top) vs νe plus ¯νe
luminosity, Lib, at the inner boundary. The symbols corre-
spond to different models of 15 M⊙ stars, the triangles to a
case including rotation (see text for details).
νe and ¯νe absorption on free nucleons.
The
"lightbulb"
approximation employed previ-
ously [6] ignored time retardation effects and did not
take into account radial variations of the fluxes due to
ν-matter interactions in the cooling and heating lay-
ers between PNS and SN shock. To improve on that
in the present work, we make use of the zeroth mo-
ment of the Boltzmann transport equation in the form
3
FIG. 2:
Two mod-
els (based on progeni-
tor WPE15) at 1 s af-
ter SN shock formation.
The PNS is at the co-
ordinate center.
The
left plot (scale reduced
to show more details)
displays a case with
Lib = 2.97×1052 erg s−1,
Eexp = 0.4 × 1051 erg,
and a recoil velocity of
vns = −350 km s−1. The
model on the right has
Lib = 4.45×1052 erg s−1,
Eexp = 1.2 × 1051 erg,
and vns = +520 km s−1.
PNS, and thus keep it in the ν-heating region to accu-
mulate energy by νe and ¯νe absorption. This is basically
in agreement with Ref. [13], where rotation was found to
stabilize the standing accretion shock at a larger radius.
The layer between PNS and SN shock is convectively
unstable according to the Ledoux-criterion because of a
negative entropy gradient established by ν-energy depo-
sition. Within ∼50 ms after shock formation, convection
sets in, supporting the start of the explosion [5, 6]. Ini-
tially the convective cells are small, but they begin to
merge to larger entities. Three-dimensional (3D) simu-
lations agree with this 2D result [19]. In case of rapid
shock acceleration convection freezes, and small struc-
tures characterize the flow pattern until late times. How-
ever, when the stagnant shock expands slowly, small cells
have time to merge to very large buoyant bubbles, sepa-
rated by only a few narrow, supersonic downflows which
carry low-entropy matter from the shock to the PNS sur-
face. Moreover, global pulsations can develop with a
dominance of low-order (l = 1, 2) modes. Consequently,
the density distribution becomes highly aspherical and
the explosion breaks out with a very large hemispheric or
polar-to-equatorial asymmetry and corresponding shock
deformation. In some cases a single long-lasting accre-
tion funnel was found to persist for a second or even
longer (Fig. 2). We emphasize that such structures did
not preferentially occur along the polar (z-) direction of
our spherical grid (where a coordinate singularity exists),
but developed in arbitrary orientations.
The development of a stable, volume-filling l = 1 mode
was proposed before [20]. It is supported by analytic ar-
guments for thermal instabilities in fluid spheres [21] and
is also observed in 3D hydrodynamic simulations of pul-
sating, convective red giant stars [22]. Determining the
duration of such a phenomenon in the time-dependent en-
vironment of an exploding SN requires numerical mod-
eling. The coherent, low-order oscillations of the fluid
beneath the shock in our simulations look similar to the
mass, Mns, for a given value of the boundary luminosity,
Lib, of νe plus ¯νe (Fig. 1). Eexp is defined as the total
energy (internal plus kinetic plus gravitational) of the
SN ejecta, integrated over all matter where the sum of
the corresponding specific energies is positive, texp is the
post-bounce time when Eexp reaches 1049 erg, and Mns
is the gas mass with densities above 1011 g cm−3 plus the
central point mass at one second.
FIG. 3: Explosion energies (bottom) and NS velocities vs time
for some simulations, showing large acceleration for cases with
high vns (thick lines) even at 1 s after bounce. The symbols
and line styles refer to the different model sequences of Fig. 1.
Figure 1 reveals the generic trend that a higher lu-
minosity Lib from the NS core causes the explosion to
develop faster and to become more energetic. Because
the period of mass accretion by the PNS is reduced,
this implies a smaller NS mass. The time until the re-
vived bounce-shock reaches a certain radius (correlated
with texp) depends sensitively on the progenitor struc-
ture and the core ν luminosity. Rotation systematically
increases the explosion energy by 20 -- 50%, because cen-
trifugal forces delay matter from being accreted onto the
recently discovered non-radial instabilities in adiabatic
flow behind standing accretion shocks [23], which can be
understood in terms of a "vortical-acoustic cycle" [24].
Doing numerical experiments we found that this phe-
nomenon might indeed play a role, although Ledoux-
instability due to ν-heating clearly starts the convective
activity, and ν-cooling around the neutrinosphere seems
to damp the energetic amplification of the feedback cycle
between turbulence and pressure waves.
Anisotropic mass ejection can be associated with a
high linear momentum taken up by the compact rem-
nant. The corresponding asymmetry of the explosion
is expressed by the parameter αgas ≡ Pz,gas/Pgas ≡
R dm vz/ R dm ~v where the integrals are performed
over the ejecta mass and Pz,gas is the gas momentum
along the z-direction of the 2D grid. We find values
for αgas up to 0.33 (Fig. 1). The NS recoil velocities,
vns = αgasPgas/Mns, can be close to zero but also more
than 500 km s−1.
In some cases the acceleration, ans,
continues on a high level beyond the 1 s of computed
evolution (Figs. 1, 3) and significantly larger terminal
velocities can be expected. It is mediated by the long-
range gravitational force between the asymmetrically dis-
tributed ejecta and the PNS. Hydrodynamic forces play
a role only as long as downflows reach the PNS, and
anisotropic ν emission contributes insignificantly. There
is no correlation of vns with the progenitor model. Rota-
tion does not inhibit large kicks. The final recoil velocity
depends stochastically on the initial seed perturbation
and the highly nonlinear growth of the convective struc-
tures. There is also no obvious correlation with the ex-
plosion energy. Fig. 3 (by the thin and thick solid lines)
and Fig. 1 (coinciding points) demonstrate that nearly
the same Eexp can be associated with large or small vns.
Conclusions. We have shown that globally anisotropic
mass ejection and NS acceleration can result from convec-
tive overturn and low-order oscillations of the ν-heated
layer in a SN core. Low-mode convection turned out to
develop from random seed perturbations in case of a slow
onset of the explosion which gives the convective struc-
tures time to merge. This was disfavored by our previ-
ous choice of a strongly time-dependent, exponentially
decaying core ν luminosity in Refs. [6, 14] but is possible
with the less burst-like (because constant) Lib assumed
in this work. Our models suggest a consistent picture in
which ν-energy deposition can be responsible for the SN
explosion, for pulsar kicks, and for global asymmetries
observed in many supernovae. The simulations need to
be continued to later times to allow for quantitative con-
clusions on the morphological properties of the ejecta as,
e.g., inferred from polarization measurements. Statisti-
cal information about the distribution of intrinsic pulsar
velocities requires 3D simulations, a better fundamental
understanding of the explosion mechanism, and a large
sample of simulations for progenitor stars with different
masses and rotation rates. The discussed kick mecha-
4
nism does not enforce a strict alignment of pulsar spin
and space velocity, although the rotation axis of a star
defines a naturally preferred direction, which might dis-
favor large misalignments.
We thank S.W. Bruenn and M. Rampp for post-bounce
core-collapse models and M. Limongi and S. Woosley
for their progenitor models. Support by the Sonder-
forschungsbereich 375 on "Astroparticle Physics" of the
Deutsche Forschungsgemeinschaft is acknowledged. The
simulations were done at the Rechenzentrum Garching
and the Interdisciplinary Centre for Computational Mod-
elling in Warsaw.
[1] A.G. Lyne and D.R. Lorimer, Nature (London) 369,
127 (1994); J.M. Cordes and D.F. Chernoff, Astro-
phys. J. 505, 315 (1998); Z. Arzoumanian, D.F. Cher-
noff, and J.M. Cordes, Astrophys. J. 568, 289 (2002);
B.M.S. Hansen and E.S. Phinney, Mon. Not. R. Astron.
Soc. 291, 569 (1997); C. Fryer, A. Burrows, and W. Benz,
Astrophys. J. 496, 333 (1998)
[2] D. Lai, D.F. Chernoff, and J.M. Cordes, Astrophys. J.
549, 1111 (2001)
[3] I.S. Shklovskii, Sov. Astron. 13, 562 (1970)
[4] A. Burrows and J. Hayes, Phys. Rev. Lett. 76, 352
(1996); D. Lai and P. Goldreich, Astrophys. J. 535, 402
(2000)
[5] M. Herant, W. Benz, W.R. Hix, C.L. Fryer, and S.A.
Colgate, Astrophys. J. 435, 339 (1994); A. Burrows, J.
Hayes, and B.A. Fryxell, Astrophys. J. 450, 830 (1995)
[6] H.-T. Janka and E. Muller, Astron. Astrophys. 290, 496
(1994); 306, 167 (1996)
[7] R. Cen, Astrophys. J. 507, L131 (1998); A.M. Khokhlov
et al., Astrophys. J. Lett. 524, L107 (1999)
[8] N.N. Chugaı, Sov. Astron. Lett. 10, 87 (1984); A. Bur-
rows and S.E. Woosley, Astrophys. J. 308, 680 (1986)
[9] G.S. Bisnovatyi-Kogan, Astron. Astrophys. Trans. 3, 287
(1993); D. Lai and Y.-Z. Qian, Astrophys. J. 505, 844
(1998)
[10] C.J. Horowitz and G. Li, Phys. Rev. Lett. 80, 3694
(1998); P. Arras and D. Lai, Astrophys. J. 519, 745
(1999); Phys. Rev. D 60, 043001 (1999)
[11] A. Kusenko and G. Segr`e, Phys. Rev. Lett. 77, 4872
(1996); E. Nardi and J.I. Zuluaga, Astrophys. J. 549,
1076 (2001)
[12] H.A. Bethe and J.R. Wilson, Astrophys. J. 295, 14
(1985)
[13] R. Buras, M. Rampp, H.-T. Janka, and K. Kifonidis,
Phys. Rev. Lett. 90, 241101 (2003)
[14] K. Kifonidis, T. Plewa, H.-T. Janka, and E. Muller, As-
tron. Astrophys. 408, 621 (2003)
[15] S.E. Woosley, P.A. Pinto, and L. Ensman, Astrophys.
J. 324, 466 (1988); S.W. Bruenn, in Nuclear Physics in
the Universe, edited by M.W. Guidry and M.R. Strayer
(IOP, Bristol, 1993), p. 31
[16] M. Limongi, O. Straniero, and A. Chieffi, Astrophys. J.
Suppl. 129, 625 (2000)
[17] S.E. Woosley and T.A. Weaver, Astrophys. J. Suppl. 101,
181 (1995)
[18] A. Heger, S.E. Woosley, N. Langer, and H.C. Spruit,
astro-ph/0301374
[19] C.L. Fryer and M.S. Warren, Astrophys. J. Lett. 574,
L65 (2002)
[22] P.R. Woodward, D.H. Porter, and M. Jacobs, in 3D Stel-
lar Evolution, ASP Conf. Series, Vol. 293 (ASP, San
Francisco, 2003), p. 45
[20] M. Herant, Phys. Rep. 256, 117 (1995); C. Thompson,
[23] J.M. Blondin, A. Mezzacappa, and C. DeMarino, Astro-
Astrophys. J. 534, 915 (2000)
phys. J. 584, 971 (2003)
[21] S. Chandrasekhar, Hydrodynamic and Hydromagnetic In-
[24] T. Foglizzo, Astron. Astrophys. 392, 353 (2002)
stabilities (Dover, New York, 1981), p. 234
5
|
astro-ph/0201302 | 1 | 0201 | 2002-01-18T13:18:26 | Observations of the unusual counterpart to the X-ray pulsar AX J0051-733 | [
"astro-ph"
] | We report optical and IR observations of the ASCA X-ray pulsar system AX J0051-733. The relationship between the X-ray source and possible optical counterparts is discussed. Long term optical data from over 7 years are presented which reveal both a 1.4d modulation and an unusually rapid change in this possible binary period. Various models are discussed. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 27 October 2018
(MN LATEX style file v1.4)
Observations of the unusual counterpart to the X-ray
pulsar AX J0051−733 ⋆
M.J. Coe1, N.J. Haigh1, S.G.T. Laycock1, I. Negueruela2 & C.R. Kaiser1
1 Physics and Astronomy Dept., The University, Southampton, SO17 1BJ, UK.
2Observatoire de Strasbourg, 11 rue de l'Universite, Strasbourg 67000, France.
Accepted
Received : Version Nov 6 2001
In original form ..
ABSTRACT
We report optical and IR observations of the ASCA X-ray pulsar system AX
J0051-733. The relationship between the X-ray source and possible optical counter-
parts is discussed. Long term optical data from over 7 years are presented which reveal
both a 1.4d modulation and an unusually rapid change in this possible binary period.
Various models are discussed.
Key words: stars: emission-line, Be - star: binaries - infrared: stars - X-rays: stars -
stars: pulsars
1
INTRODUCTION
High Mass X-ray binaries (HMXBs) are traditionally di-
vided into Be/X-ray and supergiant binary systems. A sur-
vey of the literature reveals that of the 96 proposed massive
X-ray binary pulsar systems, 67% of the identified systems
fall within the Be/X-ray group of binaries. The orbit of the
Be star and the compact object, a neutron star, is gener-
ally wide and eccentric. The optical star exhibits Hα line
emission and continuum free-free emission (revealed as ex-
cess flux in the IR) from a disk of circumstellar gas. Most of
the Be/X-ray sources are also very transient in the emission
of X-rays.
The source that is the subject of this paper, AX J0051-
733, lies in the Small Magellanic Cloud, a region of space
that is extremely rich in HMXBs. It was reported as a 323s
pulsar by Yokogawa & Koyama (1998) and Imanishi et al
(1999). Subsequently Cook (1998) identified a 0.7d optically
variable object within the ASCA X-ray error circle. The sys-
tem was discussed in the context of it being a normal HMXB
by Coe & Orosz (2000) who presented some early OGLE
data on the object identified by Cook (1998) and modelled
the system parameters. Coe & Orosz identified several prob-
lems with understanding this system, primarily that if it was
a binary then its true period would be 1.4d and it would be
an extremely compact system. In addition, the combination
⋆ Partially based on observations collected at the South African
Astronomical Observatory and the European Southern Observa-
tory, Chile (ESO 65.H-0314)
c(cid:13) 0000 RAS
of the pulse period and such a binary period violates the
Corbet relationship for such systems (Corbet, 1986).
In this paper we report on extensive new data sets from
both OGLE and MACHO, as well as a detailed photometric
study of the field. The results reveal many complex obser-
vational features that are hard to explain in the traditional
Be/X-ray binary model.
2 X-RAY SOURCE LOCATION
As will be seen from the photometric results presented be-
low, it is critical to establish the correct optical counterpart
to the X-ray pulsar. In particular, it is vital to clearly link
the ASCA source to an optical object, and other ROSAT
X-ray sources may, or may not be relevant (because no pul-
sations have been detected from ROSAT objects). Figure 1
illustrates the somewhat complex situation associated with
this object. In this figure the large dotted circle indicates the
original ASCA X-ray position and uncertainty from Imanishi
et al (1999). Within this error circle lie the much smaller er-
ror circles of the ROSAT sources RX J0050.8-7316 (Cowley
et al, 1997) and RX J0050.7-7316 (Kahabka 1998). Subse-
quently, the position of the ASCA error circle was refined
to the large solid circle shown in the figure (Imanishi 2001,
private communication).
Within the ROSAT error circle for RX J0050.7-7316 and
the revised ASCA circle for AX J0051-733 lies an obvious
optical object that has been proposed as the counterpart to
both of these X-ray objects (Cowley et al, 1997, Schmidtke
2 M.J.Coe et al.
Figure 1. Finding charts for AX J0051-733 covering a field of 4
x 4 arcmin created using an optical V band image from this work.
The northern ROSAT circle refers to RX J0050.8-7316 and the
southern one is that of RX J0050.7-7316.
& Cowley, 1998, Coe & Orosz, 2000). It is a blue star ex-
hibiting variability which strongly suggests that it is a Be
star companion to the X-ray pulsar. This object also shows a
strong 0.7d optical modulation (or possibly twice that value)
which could be associated with a binary period of the system
(Cook 1998, Coe & Orosz 2000). However, this period is very
short for a HMXB and the modulation signature atypical of
that seen from such objects.
Consequently, it was felt necessary to revisit the linking
of this optical object with the ASCA pulsar to make sure
that some other candidate was not more appropiate within
the X-ray error circle.
3 OPTICAL & IR COUNTERPART SEARCH
Optical photometric observations were taken from the
SAAO 1.0m telescope on 2 October 1996. The data were
collected using the Tek8 CCD giving a field of ∼6 x 6 ar-
cminutes and a pixel scale of 0.6 arcsec/pixel. Observations
were made through standard Johnson V & R filters plus an
Hα filter. The standard star E950 was used for photometric
calibration. From these CCD frames a R-Hα colour index
was created and this was plotted against the V band flux
for ∼800 objects.
On the assumption that our optical counterpart was
likely to be a Hα bright system, all the objects in the top
third of the colour-magnitude plot were examined and their
location in the field identified. Only four such objects were
determined to be in, or close to, the ASCA error circle. These
are numbered 476, 499, 512 and 647 in Figure 1 (object no:
512 is the proposed counterpart to the ROSAT sources). All
the other objects with an R-Hα index ≥ −1.0 lie well away
from the region of interest.
The average B, V &I colours of these four objects were
Table 1. Optical photometric values taken from the OGLE
database and IR values from the 2MASS survey.
ID
V
B-V
V-I
476
499
512
647
18.70
17.21
15.44
15.69
1.00
-0.08
-0.03
0.07
1.08
0.15
0.17
0.26
J
-
-
K
-
-
15.3
16.5
14.8
15.9
extracted from the OGLE database and are presented in
Table 1. In addition, IR magnitudes for two of the objects
are also presented that were extracted from the 2MASS sur-
vey data base, the other 2 candidates were too faint to be
detected in that survey.
To confirm the nature of Object 512 as a B or Be star,
optical spectra were obtained on 3 occasions (1 Nov 1999, 15
Sep 2000 and 22 Oct 2000) from the ESO 1.52-m telescope
at La Silla Observatory, Chile, equipped with the Boller &
Chivens spectrograph. The no: 33 holographic grating was
used, which gives a resolution of ∼1A/pixel. Since no ob-
vious variations were seen between the spectra they were
combined to increase the signal-to-noise ratio. The resulting
spectrum is presented in Figure 2. In this figure our spec-
trum is compared to that of the B0.5V standard 40 Per.
Object 512 is obviously a Be star, with Hβ and Hγ in emis-
sion and most other lines affected by emission components.
The presence of weak He ii λ4686 A places the object close
to B0V (Walborn & Fitzpatrick 1990). Though several O ii
lines are present, C iv λ4650 A is surprisingly absent. The
relatively weak Si iii and Si iv lines seen in 40 Per are not
easily detectable in object 512, which is compatible with the
lower metallicity of the SMC, but unexpected in view of the
rather strong O ii lines.
4 OGLE AND MACHO DATA
The field of AX J0051-733 lies within the areas covered by
both the OGLE ⋆ and MACHO † monitoring programmes.
Hence excellent photometric coverage exists for the brighter
counterparts for a total of nearly 7 years.
Detailed I band photometry was obtained from the
OGLE data base for objects numbered 499 (no significant
variability), 647 (some evidence for long term changes com-
parable to the length of the data set) and 512. As Cook
(1998) and Coe & Orosz (2000) have already shown from
subsets of the OGLE/MACHO data, this object exhibits a
strong clear sinusoidal modulation at ∼ 0.7d. The combined
OGLE and MACHO data set for this object is presented in
Figure 3.
Though the precise modulation is not obvious from this
figure, it clearly shows the varying amplitude of the modu-
lation over the data set. If the total data set is analysed for
periodic behaviour, then a period of 0.70872d is determined
using a Lomb-Scargle analysis. However, this period is the
average of the data, because if one splits up the data set into
∼ 150d samples a slightly different period is found for each
one.
⋆ http://sirius.astrouw.edu.pl/ ogle
† http://wwwmacho.mcmaster.ca
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
AX J0051−733
3
Figure 2. Blue spectrum of Object 512 (upper spectrum) compared to a B0.5V standard at a similar resolution. Note the presence of
relatively strong Na ii λ3934A in the spectrum of Object 512, presumably of interstellar origin.
Figure 3. Approximately 7 years of photometric observations
of the proposed counterpart to AX J0051-733 taken from the
MACHO and OGLE data bases. The date axis has MJD = JD -
2450000. In both cases the magnitude scale is indicated, though
the MACHO one is described as "approximately R".
Figure 4 illustrates the Lomb-Scargle power spectrum
for one such subset of data. To check on the aliasing with the
Nyquist frequency and the effects of the window function, a
simulated data set was created. This data set consists of a
single sinewave with period and amplitude determined from
the original data sampled with exactly the same temporal
structure as the original data. As can be seen by compari-
son between the two power spectra in Figure 4, there is no
significant difference. Thus the conclusion is that there are
no other frequencies present in the original data set.
The shape of the modulation was determined by folding
one of the MACHO and OGLE data sets at the determined
period for that data set. The result of this is illustrated in
Figure 5. Lightcurves from four different filters are shown in
this figure. In the case of the V band, the OGLE data are
rather sparse since this is not their main filter, and so the
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 4. Comparison of a Lomb-Scargle power spectrum for a
∼150d section of MACHO data (lower panel) and a simulated
data set (upper panel). The simulated data set consists of a pure
sine wave with the same window function as the raw data (see
text for more details).
4 M.J.Coe et al.
Figure 6. The lower curve shows the period history determined
from the combined MACHO and OGLE data sets. If the source is
binary system then we should expect the true binary period to be
twice the value indicated on the left hand axis. The upper curve
shows the amplitude of a sine wave fitted to each data block. In
both cases a typical error bar is indicated. The time axis has MJD
= JD - 2449000.
amount over the 7 years. The simplest interpretation of the
period change is a linear one, and from such an assumption
a value of 13.5s/year is found. More complex changes, for
example a sinusoidal modulation with a period of ∼3600d,
may be speculated upon but are not required by the data as
yet.
The upper curve in Figure 6 shows the result of fitting
a sine wave to each individual data set and determining
its amplitude. Though there is clear evidence of amplitude
variation by ∼40% from both this figure and Figure 3, it
is not obvious that there is any significant pattern to the
change.
5 DISCUSSION
5.1 Optical candidate
In trying to establish the optical counterpart to the ASCA
pulsar one must keep in mind that the ROSAT source is
too weak to have shown any detectable pulsations. Thus
putting the ROSAT source aside for the moment, the most
objective approach is to just look at the colour-magnitude
diagram (Figure ??). This diagram reveals just two objects
inside the best ASCA positional error circle - nos. 512 and
499. The other two objects lie in, or very close to the orig-
inal ASCA circle, but are now significantly less attractive
as counterparts. Object 499 is very faint compared to all
other known counterparts to HMXBs in the SMC, which
typically have V ∼ 15 − 16. Its colours and the presence of
Hα in emission suggest a B3-4Ve star - a somewhat later
spectral type than most other Be/X-ray binary systems. Its
OGLE lightcurve reveals nothing of interest and it is not a
detectable IR source in the 2MASS data. Hence it cannot be
a strong contender for the counterpart to AX J0051−733.
On the other hand, Object 512 has V = 15.4 and a
significant IR flux at J = 15.3. Both of these make it look
like a classic counterpart to a Be/X-ray binary system. If we
compare this object to another SMC X-ray pulsar system,
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 5. The lightcurves obtained by folding a ∼150d sample of
MACHO and OGLE data at the period of 1.4174d. Because the
OGLE V filter coverage is very sparse, data from several obser-
vations at SAAO in this band have been added to the illustrated
data set. The magnitude scale on the left only refers to the OGLE
I band data, all the other photometric bands have been arbitar-
ily shifted upwards by a constant amount to fit conveniently on
the figure. In each case the data sets have been phase shifted
to coalign with the OGLE I band data set (see text for further
details). The uppermost curve shows the colour information ob-
tained from the same MACHO data set used to construct the
light curve in the figure.
data have been supplemented by observations taken over
several nights from the SAAO 1.0m in January 1999 and
2000. Since this data set is not as uniform as the other 3
bands the individual results are presented in the phase di-
agram rather than a folded light curve. Overall, from this
figure, the extremely sinusoidal nature of the modulation is
very clear.
The topmost curve in Figure 5 is a measure of the colour
of the object obtained by subtracting individual MACHO
blue measurments from their red measurements taken on
the same night. If the colour data are also subjected to the
same Lomb-Scargle analysis, then the same period emerges
from the data, but the depth of modulation is clearly very
small.
Perhaps the most important result to emerge from the
combined OGLE/MACHO data set is the period history.
Figure 6 illustrates this by showing the periods determined
from the 12 individual data sets. From this figure one can
clearly see that the period is changing by a significant
1WGA J0053.8−7226 (Buckley et al, 2001), we find it is
extremely similar. In 1WGA J0053.8−7226 we have (B −
V ) = −0.06 compared to −0.03 in Object 512, and (J −
K) = 0.62 compared to 0.51 in Object 512. The E(B − V )
value found for many other SMC counterparts to Be/X-ray
systems is ∼ 0.25 (a combination of extinction to the SMC
plus local extinction due to cirumstellar material). Applying
this to the observed values for Object 512 given in Table
1 leads to an identification for the spectral type of B0III-
V. Thus even before one considers the ROSAT source, one
is led inexorably to Object 512 being the prime candidate
for the optical counterpart to AX J0051-733. The presence
of a convincing ROSAT source at the same position adds
significant extra weight to this conclusion.
The optical spectrum of Object 512 presented in Fig-
ure 2 is no later and perhaps slightly earlier than the com-
parison standard. From the colours presented in Table 1 and
assuming (B − V )0 = −0.26 (Wegner 1994), this results in
an extinction value of E(B − V ) = 0.23, which confirms
the number used above in interpreting just the photometry.
Assuming standard reddening, AV = 0.71 and therefore, as-
suming a distance modulus to the SMC (M − m) = 18.9,
the absolute magnitude for Object 512 is MV = −4.2, which
is in rather good agreement with a spectral type in the B0-
B0.5V range.
5.2 Optical modulation
The strong sinusoidal optical modulation of Object 512 is
challenging to interpret in terms of a traditional Be/X-ray
binary model. Firstly, the expected binary period of AX
J0051−733 based on the Corbet diagram (Corbet, 1986) is
100 -- 200d. Secondly, a binary period of just 1.4d involving a
Be star implies an extremely tight orbit -- the Keplerian or-
bital radius would be ∼ 14 solar radii and the B0 star has a
size of ∼ 8 solar radii. Thirdly, if the period is really decreas-
ing at a rate of 13.5 s/year then this implies (Huang 1963) a
mass transfer of 10−5 M⊙/year for mass transfer between an
18M⊙ Be star and a 1.4M⊙ neutron star -- which is not only
much larger than that typically observed in HMXB systems
(<∼ 10−8 M⊙/year in most cases), but would also imply a
much higher X-ray luminosity unless the accretion mecha-
nism is extremely inefficient at converting gravitational po-
tential into X-rays.
Mass transfer rates of this magnitude are deduced to
exist in the EB binary system β Lyrae which is changing
its ∼ 13d orbital period at a rate of 19s/year (van Hamme,
Wilson & Guinan 1995). In this case the change is to a
longer period with the mass transferring from the smaller
B6-8 star to the more massive Be star. In our case, the mass
would be flowing in the opposite direction, i.e. from the more
massive object to a less massive one. The optical lightcurve
of β Lyrae is similar to the one presented here for Object
512, but with the notable difference that in β Lyrae the two
minima are not of the same depth.
In fact the symmetry of the light curve is much more
suggestive of a W UMa type system. Unfortunately, the ob-
served period of 1.4d is much greater than any such reported
system in the SMC (Rucinski 1997). The maximum observed
period is 0.8d and our period is well off the end of the dis-
tribution. In addition, it is perhaps worth noting that the
predicted (V − I) colour obtained from the distribution of
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
AX J0051−733
5
such objects and our binary period of 1.4d is +0.026, but
from Table 1 it can be seen that the observed (V − I) for
Object 512 is 0.17. Even allowing for interstellar extinction
this further adds to it being unlikely that this system is of
this class.
The possibility of a blended variable star plus Be star
can be considered. For example, a chance superposition of
Be star (to give the observed colours) plus Cepheid or RR
Lyrae (to give the optical modulation). However, all of these
models can be ruled out because of either the magnitude of
the period, or the depth of modulation, or the shape of the
lightcurve.
Interestingly the optical modulation is somewhat simi-
lar to the short periodic modulation seen by Balona (1992)
in Be stars in the cluster NGC330 in the SMC. In this case
Balona attributes this modulation to surface features on the
rapidly rotating objects. However, how the period of such
objects could change on a timescale of years is not clear,
unless the star is in a very wide binary system. It is possible
that the data in Figure 6 could be fitted to ∼10 year sinu-
soidal modulation, but then the orbit of the neutron star
would be so distant from the Be star that it hard to see
how accretion could ever occur. In addition X-ray outbursts
have been detected 3 times over 2 years from this system
(Laycock, private communication) making such a long orbit
unlikely. Perhaps further optical data may clarify exactly
what the shape of the period change is on such timescales.
5.3 A triple system?
We are left with no convincing traditional scenario to ex-
plain all the observational data. It is very hard to see how
the orbital period change seen in Figure 6 could possibly
be caused by mass loss from a normal B0 star at a rate of
10−5 M⊙/year. One other possibility perhaps worth consid-
ering is that AX J0051−733 is a triple system -- Be star plus
another star in a tight 1.4d orbit, and the neutron star in
a highly eccentric 100−200d orbit around the pair. Such a
system could not only be intrinsically very stable since most
of the mass is concentrated in the inner binary pair, but the
transfer of angular momentum from the inner binary to the
orbit of the neutron star might also explain the evolution of
the orbital period.
o
o
Eggleton & Kiseleva (1995) derive a critical parameter
X min
for a stable triple system, which is the period ratio be-
tween the outer and inner orbits. For a system to be stable
it is required that the ratio of the orbital periods be greater
than X min
. If we assume that our inner 1.4d binary con-
sists of a B0V star (M=18M⊙) and a 1M⊙ star, while the
third outer body is a 1.4M⊙ neutron star, then this param-
eter X min
= 17. Assuming that the outer orbital period is
actually given by the position of AX J0051−733 on the Cor-
bet diagram and has a value of ∼100d, then this criterion is
easily satisfied.
o
Bailyn & Grindlay (1987) provide a formula for the rate
of change of size of the major axis of such a tight binary
(their Equation 7). Using their relationship, and assuming
one of the binary partners is the observed BO star, then it
is possible to predict the rate of change of orbital period
as a function of the mass of the other star in the inner bi-
nary. For masses of the order 15-20M⊙ the predicted period
change is ∼10 s/year. This number is in good agreement
6 M.J.Coe et al.
with the observed value of 13s/year and suggests that the
inner binary may, in fact, consist of two very similar B-
type stars. This, of course, would not present any problems
to the observed photometric or spectroscopic parameters of
the system. Even assuming that the two stars contribute
equally to the luminosity of the system would mean that
their intrinsic magnitudes are MV = −3.5, still compatible
with B0.5Ve. Furthermore, the Eggleton & Kiseleva crite-
rion remains comfortably satisfied for such a system. The
main problem raised by such scenario though would be the
very little space left between the two stars for a Be disk.
However, the evolution of this system would have to
have been very different from a classic Be/X-ray binary sys-
tem. In particular, the neutron star progenitor has probably
evolved without any mass-transfer to either of the objects
in the inner binary. Consequently it must have been much
more massive in order to have reached its current state so
long ahead of the other stars in the system. Clearly this
solution for AX J0051−733 is also not without challenges.
6 CONCLUSIONS
Detailed optical observations and analysis have been car-
ried out of the proposed counterpart to AX J0051-733. The
most likely counterpart has been identified on the basis of its
colours and Hα emission. However this object is revealed to
have a strong 0.7/1.4d modulation from long-term MACHO
and OGLE observations. Furthermore this strong period is
shown to be changing at a rate of 13.5s/year. It is hard to
reconcile all these observations with the classic Be/X-ray bi-
nary model and further studies of this system are urgently
required.
ACKNOWLEDGMENTS
We are extremely grateful to Andrzej Udalski and the OGLE
team for providing us with their data and to Kem Cook for
help with the MACHO data. We are also grateful to the
very helpful staff at the SAAO for their support during these
observations. Many helpful comments from Tom Marsh and
Dave Buckley have been very much appreciated. All of the
data reduction was carried out on the Southampton Starlink
node which is funded by the PPARC. NJH and SGTL are
in receipt of a PPARC studentships.
This paper utilizes public domain data obtained by the
MACHO Project, jointly funded by the US Department of
Energy through the University of California, Lawrence Liv-
ermore National Laboratory under contract No. W-7405-
Eng-48, by the National Science Foundation through the
Center for Particle Astrophysics of the University of Cali-
fornia under cooperative agreement AST-8809616, and by
the Mount Stromlo and Siding Spring Observatory, part of
the Australian National University.
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 Foundation.
We are also grateful to the referee, Lex Kaper, for sev-
eral helpful and interesting suggestions to improve the pa-
per.
REFERENCES
Bailyn C.D. & Grindlay J.E. 1987 ApJ 312, 748.
Balona L.A. 1992 MNRAS 256, 425.
Buckley D.A.H., Coe M.J., Stevens J.B., van der heyden K., An-
gelini L., White N.E. & Giommi P. 2001 MNRAS 320, 281.
Coe M.J. & Orosz J.A. 2000 MNRAS 311, 169.
Cook K 1998 IAUC 6860.
Corbet RHD 1986 MNRAS 220, 1047.
Cowley A.P., Schmidtke P.C., McGrath T.K., Ponder A.L., Fertig
M.R., Hutchings J.B. and Crampton D. 1997 PASP 109, 21.
Eggleton P. & Kiseleva L. 1995 ApJ 455, 640.
Huang S.S. 1963 ApJ 138, 471.
Imanishi K, Yokogawa J., Tsujimoto M. & Koyama K. 1999 PASJ
51, L15.
Kahabka P. 1998 IAUC 6854
Rucinski S.M. 1997 AJ 113, 407.
Schmidtke P.C. & Cowley A.P. 1998 IAUC 6880.
van Hamme W, Wilson R.E. & Guinan E.F. 1995 AJ 110, 1350
Walborn N.R. & Fitzpatrick E.L. 1990 PASP 102, 379.
Wegner W. 1994 MNRAS 279, 229
Yokogawa J. & Koyama K. 1998 IAUC 6853
This paper has been produced using the Royal Astronomical
Society/Blackwell Science LATEX style file.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0405275 | 1 | 0405 | 2004-05-14T06:21:44 | The distribution of cosmic-ray sources in the Galaxy, gamma-rays, and the gradient in the CO-to-H2 relation | [
"astro-ph"
] | We present a solution to the apparent discrepancy between the radial gradient in the diffuse Galactic gamma-ray emissivity and the distribution of supernova remnants, believed to be the sources of cosmic rays. Recent determinations of the pulsar distribution have made the discrepancy even more apparent. The problem is shown to be plausibly solved by a variation in the CO-to-N(H2) scaling factor. If this factor increases by a factor of 5-10 from the inner to the outer Galaxy, as expected from the Galactic metallicity gradient, we show that the source distribution required to match the radial gradient of gamma-ray can be reconciled with the distribution of supernova remnants as traced by current studies of pulsars. The resulting model fits the EGRET gamma-ray profiles extremely well in longitude, and reproduces the mid-latitude inner Galaxy intensities better than previous models. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. ms
(DOI: will be inserted by hand later)
June 26, 2018
The distribution of cosmic-ray sources in the Galaxy, γ-rays and
the gradient in the CO-to-H2 relation
4
0
0
2
y
a
M
4
1
1
v
5
7
2
5
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A. W. Strong1, I. V. Moskalenko2
,
3, O. Reimer4, S. Digel5, and R. Diehl1
1 Max-Planck-Institut fur extraterrestrische Physik, Postfach 1312, D-85741 Garching, Germany
2 NASA/Goddard Space Flight Center, Code 661, Greenbelt, MD 20771, USA
3 Joint Center for Astrophysics, University of Maryland, Baltimore County, Baltimore, MD 21250, USA
4 Ruhr-Universitat Bochum, D-44780 Bochum, Germany
5 W.W. Hansen Experimental Physics Laboratory, Stanford University, Stanford, CA 94305, USA
Received / Accepted
Abstract. We present a solution to the apparent discrepancy between the radial gradient in the diffuse Galactic
γ-ray emissivity and the distribution of supernova remnants, believed to be the sources of cosmic rays. Recent
determinations of the pulsar distribution have made the discrepancy even more apparent. The problem is shown to
be plausibly solved by a variation in the WCO-to-N (H2) scaling factor. If this factor increases by a factor of 5 -- 10
from the inner to the outer Galaxy, as expected from the Galactic metallicity gradient, we show that the source
distribution required to match the radial gradient of γ-rays can be reconciled with the distribution of supernova
remnants as traced by current studies of pulsars. The resulting model fits the EGRET γ-ray profiles extremely
well in longitude, and reproduces the mid-latitude inner Galaxy intensities better than previous models.
Key words. gamma rays -- Galactic structure -- interstellar medium -- cosmic rays -- supernova remnants -- pulsars
1. Introduction
The puzzle of the Galactic γ-ray gradient goes back to
the time of the COS-B satellite (Bloemen et al., 1986;
Strong et al., 1988); using HI and CO surveys to trace
the atomic and molecular gas, the Galactic distribution
of emissivity per H atom is a measure of the cosmic-ray
(CR) flux, for the gas-related bremsstrahlung and pion-
decay components. However the gradient determined in
this way is much smaller than expected if supernova rem-
nants (SNR) are the sources of cosmic rays, as is gen-
erally believed. This discrepancy was confirmed with the
much more precise data from EGRET on the COMPTON
Gamma Ray Observatory, even allowing for the fact that
inverse-Compton emission (unrelated to the gas) is more
important than originally supposed (Strong et al., 2000).
A possible explanation of the small gradient in terms of
CR propagation, involving radial variations of a Galactic
wind, was recently put forward by Breitschwerdt et al.
(2002).
However the derivation of the Galactic distribution of
SNR, commonly based on radio surveys, is subject to large
observational selection effects, so that it can be argued
that the discrepancy is not so serious. But other trac-
Send offprint requests to: A. W. Strong, [email protected]
ers of the distribution of SNR are available, in particular
pulsars; the new sensitive Parkes Multibeam survey with
914 pulsars has been used by Lorimer (2004) to derive
the Galactic distribution, and this confirms the concen-
tration to the inner Galaxy. Fig. 1 compares the pulsar
distribution from Lorimer (2004) with a CR source distri-
bution which fits the EGRET γ-ray data (Strong et al.,
2000). If the pulsar distribution indeed traces the SNR,
then there is a serious discrepancy with γ-rays. The dis-
tribution of SNR given by Case & Bhattacharya (1998) is
not so peaked, but the number of known SNR is much
less than the number of pulsars and the systematic ef-
fects very difficult to account for (Green, 1996). But even
this flatter distribution is hard to reconcile with that re-
quired for γ-rays. Another, quite independent, tracer of
the SNR distribution is the 1809 keV line of 26Al; whether
this originates mainly in type II supernovae or mass-
sive stars is not important in this context, since both
trace star-formation/SNR. The COMPTEL 26Al maps
(Knodlseder et al., 1999; Pluschke et al., 2001) show that
the emission is very concentrated to the inner radian of the
Galaxy. The density of free electrons shows a similar dis-
tribution (Cordes & Lazio, 2003). The 26Al measurements
are not subject to the selection effects of other methods;
2
Strong A.W. et al.: Distribution of cosmic ray sources in the Galaxy
Fig. 1. CR source density as function of Galactocentric
radius R. Dotted: as used in Strong et al. (2000), solid
line:based on pulsars (Lorimer, 2004) as used in this work,
vertical bars: SNR data points from Case & Bhattacharya
(1998). Distributions are normalized at R = 8.5 kpc.
although they have their own uncertainties, they support
the type of distribution which we adopt in this paper.
A major uncertainty in the models of diffuse Galactic
γ-ray emission is the distribution of molecular hydro-
gen, as traced by the integrated intensity of the J = 1 --
0 transition of 12CO, WCO. Gamma-ray analyses have
in fact provided one of the standard values for the
scaling factor1 XCO= N (H2)/WCO; with only the as-
sumption that cosmic rays penetrate molecular clouds
freely, the γ-ray values are free of the uncertainties of
other methods (e.g. those based on the assumption of
molecular cloud virialization). However previous analy-
ses, e.g. Strong & Mattox (1996), Hunter et al. (1997),
Strong et al. (2000), have usually assumed that XCO is
independent of Galactocentric radius R, since otherwise
the model has too many free parameters. But there is now
good reason to believe that XCO increases with R, both
from COBE/DIRBE studies (Sodroski et al., 1995, 1997)
and from the measurement of a Galactic metallicity gradi-
ent combined with the strong inverse dependence of XCO
on metallicity in external galaxies (Israel, 1997, 2000). A
rather rapid radial variation of XCO is expected, based on
a gradient in [O/H] of 0.04 -- 0.07 dex/kpc (Hou et al., 2000;
Deharveng et al., 2000; Rolleston et al., 2000; Smartt,
2001; Andrievsky et al., 2002) and the dependence of
XCO on metallicity in external galaxies: log XCO∝ −2.5
[O/H] (Israel, 1997, 2000), giving XCO∝ 10(−0.14±0.04)R,
amounting to a factor 1.3 -- 1.5 per kpc, or an order of
magnitude between the inner and outer Galaxy. 2 A less
rapid dependence, log XCO∝ -- 1.0 [O/H], was found by
1 units: molecules cm−2/ (K km s−1)
2 The values given by Israel (1997, 2000) include the ef-
fects of the radiation field, implicitly containing the radiation
field/metallicity correlation of his galaxy sample. XCO is posi-
tively and almost linearly correlated with radiation field, so the
dependence of XCO for constant radiation field is even larger:
log XCO∝ −4 [O/H] (Israel, 2000). By adopting the coefficient
-- 2.5 we implicitly assume the same radiation/metallicity cor-
relation within the Galaxy as over his galaxy sample.
Fig. 2. XCO as function of R. Dotted horizontal line,
black: as used in Strong & Mattox (1996); Strong et al.
(2000); solid line, black: as used for γ-rays in this work;
dashed, dark blue: from Sodroski et al. (1995); dash-dot,
red: using metallicity gradient as described in the text,
XCO∝ Z −2.5 (Israel, 2000), two lines for [O/H] = 0.04 and
0.07 dex/kpc; dash-dot-dot,light blue: using XCO∝ Z −1.0
(Boselli et al., 2002) and [O/H] = 0.07 dex/kpc. The val-
ues using metallicity are normalized approximately to
those from the γ-ray analysis.
Boselli et al. (2002), which however still implies a signifi-
cant XCO(R) variation. Boissier et al. (2003) also combine
the metallicity gradient with XCO(Z) within individual
galaxies, to obtain radial profiles of H2, and give argu-
ments for the validity of this procedure. Digel et al. (1990)
found that molecular clouds in the outer Galaxy (R∼12
kpc) are underluminous in CO, with XCO a factor 4±2
times the inner Galaxy value. Sodroski et al. (1995, 1997)
derived a similar variation (logXCO/1020 = 0.12R − 0.34)
when modelling dust emission for COBE data. Pak et al.
(1998) predicted the physical origin for a variation of XCO
with Z. Papadopoulos et al. (2002) and Papadopoulos
(2004) discuss the physical state of this metal-poor gas
phase in the outer parts of spiral galaxies (relatively
warm and diffuse). Observations of H2 line emission from
NGC 891 with ISO (Valentijn & van der Werf, 1999) in-
dicate a massive cool molecular component in the outer
regions of this galaxy, supporting the trend found in our
Galaxy.
Fig. 2 illustrates some of the possible XCO variations
implied by these studies. For the cases where metallicity is
used to estimate XCO, the values are normalized approx-
imately to the values used in the present γ-ray analysis,
since we are only interested in comparing the variations of
XCO. From the viewpoint of γ-rays, the effect of a steeper
CR source distribution is compensated by the increase of
XCO. Thus we might expect to resolve the apparent dis-
crepancy in the source distribution, and improve our un-
derstanding of the Galactic γ-ray emission. In this paper
we investigate quantitatively this possibility. Note that the
γ-rays include major contributions from interactions with
atomic hydrogen and from inverse Compton scattering,
both of which are independent of XCO; this means that
Strong A.W. et al.: Distribution of cosmic ray sources in the Galaxy
3
-410·
0.5
galdef ID 45_600202
1000 - 2000 MeV
-5.50<b<-0.50 , 0.50<b< 5.50
galdef ID 45_600202
1000 - 2000 MeV
0.50<l<30.50 , 330.50<l<359.50
-410
1
-
s
1
-
r
s
2
-
m
c
-510
,
y
t
i
1
-
s
1
-
r
s
2
-
m
c
,
y
t
i
s
n
e
t
n
i
0.4
0.3
0.2
0.1
0
180
180
180
180
s
n
e
t
n
i
-610
-710
160
160
160
160
140
140
140
140
120
120
120
120
100
100
100
100
80
80
80
80
60
60
60
60
40
40
40
40
20
20
20
20
0
0
0
0
340
340
340
340
320
320
320
320
300
300
300
300
280
280
280
280
260
260
260
260
240
240
240
240
220
220
220
220
200
200
200
200
180
180
180
180
Galactic longitude
Galactic longitude
Galactic longitude
Galactic longitude
-80 -60 -40 -20
0
20
40
60
80
Galactic latitude
Fig. 3. Longitude profile of γ-rays for 1000 -- 2000 MeV,
averaged over b < 5.5◦. Vertical bars: EGRET data; lines
are model components convolved with the EGRET point-
spread function: green: inverse Compton emission, red: π0-
decay, light blue: bremsstrahlung, dark blue: total.
Fig. 4. Latitude profile of γ-rays for 1000 -- 2000 MeV, av-
eraged over 330◦ < l < 30◦. Data and curves as in Fig. 3.
The extragalactic background is shown as a black horizon-
tal line.
the XCO variation has to be quite large to have a signifi-
cant effect.
2. Data
The EGRET and COMPTEL data are the same as de-
scribed in Strong et al. (2000, 2004a). The EGRET data
consist of the standard product counts and exposure for
30 MeV -- 10 GeV, augmented with data for 10 -- 120 GeV.
The γ-ray point sources in the 3EG catalogue have been
removed as described in Strong et al. (2000). The HI and
CO data are as described in Moskalenko et al. (2002) and
Strong et al. (2004a); they consist of combined surveys di-
vided into 8 Galactocentric rings on the basis of kinematic
information. Full details of the procedures for compar-
ing models with data are given in Strong et al. (2004a) to
which the reader is referred.
3. Model and Method
We use the GALPROP program (Strong et al., 2000,
2004a) to compute the models. GALPROP was extended
to allow a variable XCO(R) to be input. The distribu-
tion of CR sources is assumed to follow that of pulsars
in the form given by Lorimer (2004), as shown in Fig. 1.
The other parameters, in particular the CR nucleon and
electron injection spectral shape and propagation parame-
ters, are taken from the "optimized model" of Strong et al.
(2004a). As before the halo height is taken as zh = 4
kpc, and the maximum radius R = 20 kpc. The isotropic
background is as derived in Strong et al. (2004b). Since
in this work we simply wish to demonstrate the possi-
bility to obtain a plausible solution, we adopt a heuristic
approach, adjusting XCO(R) to obtain a satisfactory solu-
tion as shown in Fig. 2. The electron flux has been scaled
down by a factor 0.7 relative to Strong et al. (2004a) to
obtain an optimal fit.
4. Results
Figs. 3 and 4 show the longitude and latitude distributions
for 1 -- 2 GeV, compared to EGRET data. A rather rapid
variation of XCO is required to compensate the CR source
gradient, but it is fully compatible with the expected vari-
ation based on metallicity gradients and the COBE result,
as described in the Introduction. The longitude and lati-
tude fits are good except in the outer Galaxy where the
prediction is rather low. One possible reason for this is
that the CR source density does not fall off so fast beyond
the Solar circle as given by the adopted pulsar distribu-
tion, which has an exponential decay. Another possibility
could be even larger amounts of H2 in the outer Galaxy
than we have assumed (see discussion in Introduction).
We have chosen the range 1 -- 2 GeV for the profiles since
this is where the gas contribution and hence the effect of
XCO is maximal. An exhaustive comparison of profiles in
all energy ranges is beyond the scope of this Letter, but
in fact the agreement is good at all energies. The larger
CR gradient in this model has another consequence: the
predicted inverse-Compton emission in the inner Galaxy is
more intense at intermediate latitudes where the interstel-
lar radiation field is still high; this is precisely the region
where previous models (Hunter et al., 1997; Strong et al.,
2000, 2004a) have had problems to reproduce the EGRET
data. Fig. 5 shows the model spectrum of the inner Galaxy
compared with EGRET data; the fit is similar to that of
models (Strong et al., 2004a) with ad hoc source gradient
and constant XCO. The prediction is rather high above 20
GeV, however the EGRET data are least certain in this
range (Strong et al., 2004a).
5. Discussion
We have shown that a good fit to the EGRET data is
obtained with the particular combination of parameters
chosen. We can however ask whether the pulsar source dis-
tribution combined with a constant XCO could also give
4
V
e
M
1
-
s
1
-
r
s
2
-
m
c
,
y
t
i
s
n
e
t
n
i
.
2
E
Strong A.W. et al.: Distribution of cosmic ray sources in the Galaxy
galdef ID 45_600202
-110
0.50<l<30.50 , 330.50<l<359.50
-5.50<b<-0.50 , 0.50<b< 5.50
acknowledges support from the BMBF through DLR grant
QV0002.
total
IC
-210
-310
-410
1
References
Andrievsky, S. M., Bersier, D., Kovtyukh, V. V., et al.
2002, A&A, 384, 140
EB
Boissier, S., Prantzos, N., Boselli, A., & Gavazzi, G. 2003,
MNRAS, 346, 1215
Bloemen, J. B. G. M., Strong, A. W., Mayer-
Hasselwander, H. A., et al. 1986, A&A, 154, 25
bremss
10
210
310
410
610
energy, MeV
510
Breitschwerdt, D., Dogiel, V. A., & Volk, H. J. 2002, A&A,
Fig. 5. Spectrum of inner Galaxy, 330◦ < l < 30◦, b <
5.5◦. Vertical bars: EGRET data (red), COMPTEL data
(green). Curves: predicted intensities; inverse Compton
(green), π0-decay (red), bremsstrahlung (light blue), ex-
tragalactic background (black), total (dark blue).
a good fit if we reduce the CR electron intensity, to su-
press the inner Galaxy peak from inverse Compton emis-
sion. This can indeed reproduce the longitude profile in the
inner Galaxy, but fails badly to account for the latitude
distribution, since it has a large deficit at intermediate lat-
itudes. Some variation of XCO is therefore required. The
suggested variation of XCO would have significant impact
on the Galactic H2 mass and distribution. Warm molec-
ular hydrogen in the outer parts of spiral galaxies that
is not traced by CO emission may be detectable by the
Spitzer observatory in 28 µm vibrational emission. These
issues will be addressed in future work.
6. Conclusions
Two a priori motivated developments allow us to obtain
a more physically plausible model for Galactic γ-rays, si-
multaneously allowing a CR source distribution similar to
SNR as traced by pulsars and an expected variation in the
WCO-to-N (H2) conversion factor. Obviously the uncer-
tainty in both the source distribution and XCO are large
so our solution is far from unique, but it demonstrates the
possibility to obtain a physically-motivated model without
resorting to an ad hoc source distribution. This result sup-
ports the SNR origin of CR. The resulting model also gives
improved predictions for γ-rays in the inner Galaxy at
mid-latitudes. We have therefore achieved a step towards
a better understanding of the diffuse Galactic γ-ray emis-
sion. This result is important input to the development of
models for the upcoming GLAST mission. This Letter is
intended only to point out the potential importance of the
effect. The next step will be a more quantitative analysis
to derive XCO(R) from the γ-ray data themselves.
Acknowledgements. We thank F. Israel and D. Lorimer and the
referee for useful discussions. I.V.M. acknowledges partial sup-
port from a NASA Astrophysics Theory Program grant, O.R.
385, 216
Boselli, A., Lequeux, J., & Gavazzi, G. 2002, A&A, 384,
33
Case, G. L., & Bhattacharya, D. 1998, ApJ, 504, 761
Cordes, J. M., & Lazio, T. J. W. 2003, astro-ph/0207156
Deharveng, L., Pena, M., Caplan, J., & Costero, R. 2000,
MNRAS 311, 329
Digel, S., Bally, J., & Thaddeus, P. 1990, ApJ, 357, L29
Green, D. A. 1996,
in Supernovae and Supernova
Remnants, IAU Coll. 145, R. McCray and Z. Wang
(eds), Cambridge University Press, p.341
Hou, J. L., Prantzos, N., & Boissier, S. 2000, A&A 362,
921
Hunter, S. D., Bertsch, D. L., Dingus, B.L., et al. 1997,
ApJ, 481, 205
Knodlseder, J., Dixon, D. D., Bennett, K., et al. 1999,
A&A 344, 68
Israel, F. P. 1997, A&A, 328, 471
Israel, F. P. 2000, in Molecular Hydrogen in Space, F.
Combes and G. Pineau des Forets (eds), p.293
Lorimer, D. R. 2004, in Young Neutron Stars and Their
Environments, IAU Symp. 218, F. Camilo and B. M.
Gaensler (eds), astro-ph/0308501
Moskalenko, I. V., Strong, A. W., Ormes, J. F., &
Potgieter, M. S. 2002, ApJ, 565, 280
Papadopoulos, P. P., Thi, W.-F., & Viti, S. 2002, ApJ,
579, 270
Papadopoulos, P. P. 2004, in The Neutral ISM in Starburst
Galaxies, PASP Conf. Series, in press, astro-ph/0403087
Pak, S., Jaffe, D. T., van Dishoeck, E. F., Johansson, L.
E. B., Ewine, F. & Booth, R. S. 1998, ApJ, 498, 735
Pluschke, S., Diehl, R., Schonfelder, V., et al. 2001, ESA
SP -- 459, 55
Rolleston, W. R. J., Smartt, S. J., Dufton, P. L., & Ryans,
R. S. I. 2000, A&A, 363, 537
Sodroski, T. J., Odegard, T. J., Dwek, E., et al. 1995, ApJ,
452, 262
Sodroski, T. J., Odegard, N., Arendt, R., et al. 1997, ApJ,
480, 173
Smartt, S. J., Venn, K. A., Dufton, P. L., Lennon, D. J.,
Rolleston, W. R. J., & Keenan, F. P. 2001, A&A, 367,
86
Strong, A. W., Bloemen, J. B. G. M., Dame, T. M., et al.
1988, A&A, 207, 1
Strong, A. W., & Mattox, J. R. 1996, A&A, 308, L21
p
o
Strong A.W. et al.: Distribution of cosmic ray sources in the Galaxy
5
Strong, A. W., Moskalenko, I. V., & Reimer, O. 2000, ApJ,
537, 763
Strong, A. W., Moskalenko, I. V., & Reimer, O. 2004a,
submitted to ApJ
Strong, A. W., Moskalenko, I. V., & Reimer, O. 2004b,
submitted to ApJ
Valentijn, E. A. & van der Werf, P. P. 1999, ApJ, 522, L29
|
astro-ph/9701049 | 1 | 9701 | 1997-01-10T02:03:02 | Energy Budget and the Virial Theorem in Interstellar Clouds | [
"astro-ph"
] | The Virial Thoerem is a mathematical expression obtained from the equation of motion for a fluid, which describes the energy budget of particular regions within the flow. This course reviews the basic theory leading to the Virial Theorem, discusses its applicability and limitations, and then summarizes observational results concerning the physical and statistical properties of interstellar clouds which are normally understood in terms of the Virial Theorem, in particular the so-called ``Larson's Relations''. In all cases, the standard notions as well as relevant counterpoint are presented. In particular, the difficulties arising when the medium is fully turbulent are discussed. | astro-ph | astro-ph |
Energy Budget and the Virial Theorem in Interstellar
Clouds
Enrique V´azquez-Semadeni
Instituto de Astronom´ia, UNAM
Apartado Postal 70-264, M´exico D.F. 04510, MEXICO
Abstract. The Virial Thoerem (VT) is a mathematical expression
obtained from the equation of motion for a fluid, which describes the
energy budget of particular regions within the flow. This course re-
views the basic theory leading to the VT, discusses its applicability
and limitations, and then summarizes observational results concern-
ing the physical and statistical properties of interstellar clouds which
are normally understood in terms of the VT, in particular the so-
called "Larson's Relations". In all cases, the standard notions as well
as relevant counterpoint are presented. In particular, the difficulties
arising when the medium is fully turbulent are discussed.
To appear in "Millimetric and Sub-Millimetric Astronomy. INAOE 1996
Summer School".
1.
Introduction
The interstellar medium is an extremely complex mixture of gas, dust, and
cosmic rays, threaded by a ubiquitous magnetic field, and stirred by a
wide variety of energy sources such as supernovae, expanding HII regions
and bipolar outflows. Progressively higher-density regions in this medium
constitute what we normally call intercloud medium, cloud complexes, dif-
fuse clouds, molecular clouds, clumps and cores, respectively, although the
boundaries between these classifications are fuzzy (a natural consequence of
drawing arbitrary classification boundaries in a continuum). Table 1. gives
the ranges of densities and temperatures for the above structures (adapted
from Jura 1987).
The gaseous component of this medium should be reasonably well de-
scribed as a compressible, self-gravitating fluid (e.g, Shu 1992, Ch. 1) obey-
ing the magnetohydrodynamic (MHD) equations, since even very low ion-
ization fractions allow coupling of the motions to the magnetic field (Mestel
& Spitzer 1956; see also Shu 1992, Ch. 27). From one of these equations,
a very important result called the Virial Theorem (VT) can be obtained,
which describes the energy balance (or "budget") of particular regions within
1
2
Enrique V´azquez-Semadeni
Table 1.
(Adapted from Jura 1987.)
Structures in the interstellar medium.
Name
n(cm−3) T (K)
Filling factor
Hot intercloud 3 × 10−3
gas
Warm gas
0.1
106
0.1 − 0.7?
103 − 104
0.4
Diffuse clouds < 10
50 − 100
Cirrus clouds
10 − 103
10 − 100
Dark clouds
GMCs
H II regions
> 103
> 103
> 10
10
0.01?
15 − 40
104
SNR
> 1
104 − 107
Virial Theorem in Interstellar Clouds
3
the medium.
In this course, we first review the derivation of the VT (§
2.), stressing in particular the physical meaning of each term, including the
often-neglected surface terms, and the validity of frequently used simplifying
assumptions (e.g., Shu 1992). In § 3. we then discuss various applications
of the Virial Theorem, both to idealized cases and to real interstellar cloud
properties, pointing out existing incompletenesses. Finally, in § 4. contains
the conclusions. Throughout this course, we shall ignore the flux-loss effects
of ambipolar diffusion (e.g, Shu 1992).
2. The MHD Equations and the Virial Theorem
2.1. The MHD equations
In order for the gaseous component of the ISM to be adequately described
by a continuum approximation, it is necessary that the mean free paths of
the gas molecules be much smaller than the characteristic length scale of the
systems under consideration (e.g., Currie 1974). This is generally satisfied
in most kinds of interstellar structures, since the characteristic length scales
are normally at least some 105 times larger than the mean free paths (e.g.,
Shu 1992). The use of the MHD equations is thus amply justified. These
equations, describing the evolution of a magnetized, compressible gas with-
out dissipative terms, are (e.g., Cowling 1976; Spitzer 1978; Shu 1992; Shore
1992):
i) The Continuity Equation, expressing mass conservation:
∂ρ
∂t
+ ∇ · (ρu) = 0;
(1)
ii) The Momentum Balance Equation
ρ∂u
∂t
+ ρu · ∇u = −∇P − ρ∇ϕ +
1
4π
(∇ × B) × B − 2ρΩ × u;
(2)
iii) The Internal Energy Balance Equation
∂e
∂t
+ u · ∇e = −(γ − 1)e∇ · u + Γ − ρΛ;
iv) The Magnetic Field Equation (flux freezing)
v) Poisson's Equation
∂B
∂t
= ∇ × (u × B);
∇2ϕ = 4πGρ.
(3)
(4)
(5)
Additionally, we assume an ideal-gas equation of state P = (γ − 1)ρe.
4
Enrique V´azquez-Semadeni
In these equations, ρ is the mass density of the gas, u is the fluid velocity,
P is the thermal pressure, e is the specific internal energy, B is the magnetic
field, and ϕ is the gravitational potential. Additionally, Γ and Λ symbolically
represent any heating and cooling sources present in the medium, and are
normally functions of the density and temperature in the MHD description
(neglecting radiative transfer and chemical composition).
Let us now briefly discuss the meaning of the main terms in these equa-
tions. In eq. (2), which is Newton's Second Law in per-unit-volume form,
the second term on the left hand side (LHS) is the so-called non-linear or
advective term, and expresses the momentum transport by the velocity field
itself. This is the term that describes the effects of turbulence in the medium.
The first term in the right hand side (RHS) gives the force (per unit vol-
ume) exerted by the thermal pressure gradient. The second term is the
self-gravitational force, the third term is the Lorentz force, and the last
term is the Coriolis force, due to Galactic rotation with an angular velocity
Ω.
In eq. (3), the second term on the LHS expresses internal energy trans-
port by the velocity field (convection). The first term on the RHS is the P dV
work. In eq. (4), the RHS is the flux freezing condition, implying that mag-
netic field lines are dragged along with the matter. Finally, eq. (5) expresses
that the source of the gravitational potential is the mass density.
Important physics left out from this description are, as mentioned above,
the radiative transfer and the chemical and ionization fraction evolution.
These can be crudely incorporated in the equations above by means of model
terms, such as Γ and Λ in eq. (3), or by making the equation of state vary
in different regimes. This allows an approximate description of the effects of
these processes on the dynamics, although it cannot allow any predictions
on those processes themselves. Finally, note that the MHD equations hold
in the so-called "MHD approximation" which assumes an conductivity of
the medium.
2.2. The Virial Theorem
The scalar VT is obtained by dotting the momentum equation (eq. 2) with
the position vector x and integrating over some volume V . Traditionally, the
volume V is taken in a Lagrangian frame of reference moving with the fluid
(e.g., Spitzer 1978; Shu 1992). An Eulerian (i.e, in a fixed frame of reference)
version of the VT has recently been derived by McKee & Zweibel (1992,
hereafter MZ92). This may be a more convenient form for computation in
practice (e.g., MZ92; Ballesteros-Paredes & V´azquez-Semadeni 1997), but
here we discuss the Lagrangian version for consistency with most treatments.
Neglecting the Coriolis force, which is negligible at molecular cloud scales,
Virial Theorem in Interstellar Clouds
5
one obtains
ZV
ρx·
du
dt
dV = −ZV
x·∇P dV −ZV
ρx·∇ϕdV +
1
4π ZV
x·(∇×B)×BdV, (6)
where d/dt = ∂/∂t + u · ∇ is the total or Lagrangian derivative operator.
The LHS can be rewritten as
ZV
ρx ·
du
dt
dV =
1
2 Z d2x2
dt2 dm − ZV
ρu2dV ≡
1
2
d2I
dt2 − ZV
ρu2dV,
(7)
where we have defined dm = ρdV , and have used Reynolds' Transport The-
orem
d
dt ZV
αdV = ZV h ∂α
dt dm. Also, the second equality in eq. (7) defines
+ ∇ · (αu)idV
(8)
∂t
to show that d
the moment of inertia I.
dt R αdm = R dα
For the RHS, consider the following. First, according to Newton's Law
of Gravitation (or from the solution of Poisson's equation) the gravitational
term can be written as
W ≡ −ZV
ρ(x)x · ∇ϕdV = −GZV Zall space
ρ(x)ρ(x′)x · (x − x′)
x − x′3
dV ′dV. (9)
Noting that the integrand is nearly symmetric with respect to the primed
and unprimed variables, we can replace x by (x − x′)/2 and write
W = −
1
2
GZV Zall space
ρ(x)ρ(x′)
x − x′
dV dV ′.
(10)
Thus, W is the gravitational energy of the mass distribution within volume
V , given the mass distribution of the whole space.1 Note that in Shu (1992),
both integrals are evaluated over volume V . This approximation is valid
only if the mass outside volume V can be neglected (see Spitzer 1978), a
fact which is often overlooked.
Secondly, the remaining thermal pressure and magnetic terms can be
dealt with using Gauss' divergence theorem, and the facts that ∇ · x = 3
and that (in tensor notation) ∂xi/∂xj = δij. Additionally, for the magnetic
term, it is most convenient to work with Maxwell's magnetic stress tensor
1Interestingly, this is not true in two dimensions (2D) (Ballesteros-Paredes & V´azquez-
Semadeni, in preparation), due to the different spatial variation of the gravitational po-
tential, analogously to the well-known situation for electromagnetic fields. This is un-
fortunate, since many numerical simulations are performed in 2D due to computational
limitations.
6
Enrique V´azquez-Semadeni
T ≡ Tij ≡ BiBj/4π − B2δij/8π, such that the flux freezing term satisfies
(∇ × B) × B)/4π = ∇ · Tij. Thus, the VT finally reads
d2I
1
2
dt2 = Z ρu2dV + (cid:16)3Z P dV − IS
+(cid:16) 1
x · T · dS(cid:17)
P x · dS(cid:17)
−
8π Z B2dV + IS
GZV Zall space
≡ 2K + (cid:16)2U − IS
1
2
ρ(x)ρ(x′)
x − x′
dV dV ′
P x · dS(cid:17) + (cid:16)M + IS
x · T · dS(cid:17) + W,
(11)
where S is the surface enclosing V .
Several comments and warnings are in order. First, note that the LHS is
the second time derivative of the moment of inertia of the cloud. The cloud
is said to be in virial equilibrium if the RHS is zero. However, note that in
particular this includes the case of a constant rate of change in the cloud's
moment of inertia. Thus, virial equilibrium does not necessarily imply a
stationary or static cloud configuration, although the expression is almost
always interpreted to mean so.
Second, note that both the thermal pressure and the magnetic terms
contain a "volumetric" and a "surface" contribution. In fact, if either the
thermal pressure or the magnetic field are uniform, then the corresponding
surface and volumetric terms cancel each other, indicating balance between
internal and external stresses. Conversely, the surface terms can only be
ignored if the corresponding fields can be neglected at the boundary surface.
However, it is frequently encountered in the literature that virial balance is
considered only among the volumetric terms.
Third, the thermal pressure surface term describes the "confining" ef-
fect of the external pressure on the cloud (volume V ). A similar comment
applies to the magnetic surface term, although in this case this term is not
exclusively confining, as it includes the effects of both the external mag-
netic pressure as well as the magnetic tension along field lines. Moreover, in
the Eulerian version of the theorem, an analogous surface term of the form
R x · ρuu · dS for the macroscopic velocity field also appears (MZ92), where
ρuu is a momentum transfer tensor reminiscent of the Reynolds' stress ten-
sor (see, e.g., Landau & Lifshitz 1987, § 42; Lesieur 1990, Ch. VI). However,
due to the strongly fluctuating nature of the turbulent velocity field, which
may contain large-scale motions (comparable to the cloud's size), this term
can hardly be considered as producing a confining effect; instead, it produces
transport of momentum across the cloud's boundary and a redistribution of
the mass within the cloud, thus contributing to the change in its moment of
inertia. Neglect of this type of exchange across an Eulerian boundary, or of
the high mobility of a Lagrangian boundary, led MZ92 to the conclusion that
Virial Theorem in Interstellar Clouds
7
the intercloud velocity component perpendicular to the surface of the cloud
must be small at small distances from the boundary surface S, a result which
appears questionable in the light of numerical simulations of turbulence in
the ISM (Passot et al. 1988; L´eorat et al 1990; V´azquez-Semadeni et al.
1995; Passot et al. 1995), which suggest an extremely dynamical scenario.
The last point leads to a practical warning. The choice of the vol-
ume V and its boundary S in the actual ISM or in numerically simulated
flows constitutes a difficult task as soon as high enough resolution is avail-
able. The VT has traditionally been applied to idealized spheroidal, static
clouds either in magnetohydrodynamic equilibrium (possibly assisted by ex-
ternal pressure confinement; e.g., Strittmatter 1966; Mouschovias & Spitzer
1976; see also Shu 1992), equilibrium between internal (micro)turbulence and
self-gravity (Chandrasekhar 1951; Bonazzola et al. 1987; V´azquez-Semadeni
& Gazol 1995), or combinations thereof (e.g, Myers & Goodman 1988a,b;
Mouschovias & Psaltis 1995). In these cases, the clouds have well defined
boundaries. However, if the clouds are highly dynamical, even possibly with
fractal structure (Scalo 1990; Falgarone et al. 1991), and they are observed
with high enough resolutions (e.g., Bally et al. 1987; see also the numerical
results of Ballesteros-Paredes & V´azquez-Semadeni 1997), then important
difficulties arise. In a Lagrangian description, the cloud boundaries (the sur-
faces S) move in an extremely complex way, getting stretched and distorted.
In an Eulerian description, a surface S defined at a particular time will
have an extremely amorphous shape, possibly being very elongated or hav-
ing "tentacle"-like protrusions. Thus, at subsequent times, mass will have
entered or left the contained volume V , rendering the identity of the cloud
dubious (Ballesteros-Paredes & V´azquez-Semadeni 1997). In this case, it is
likely that all the terms in the RHS of the VT have comparable importance,
particularly the surface terms. With this caveat in mind, we now proceed to
review some of the best known applications of the VT to interstellar clouds.
3. Applications
3.1.
Ideal cases
Pressure confinement. Taking B = u = 0 and neglecting self-gravity in
eq. (11), one gets the condition of virial equilibrium, d2I/dt2 = 0 ⇒ 2U =
HS P x · dS, indicating balance between internal and external pressure. This
is a good example of a case in which volume V can be chosen in a completely
arbitrary way, since this is the case, for example, of any parcel of gas in the
air in the room you are in.
Hydrostatic equilibrium and turbulent support. Assuming B = u = 0 and
a vanishing thermal pressure outside volume V , but keeping self-gravity, we
obtain d2I/dt2 = 0 ⇒ W = −2U , indicating hydrostatic equilibrium. In-
8
Enrique V´azquez-Semadeni
cluding a turbulent velocity field, the condition of virial equilibrium becomes
W = −2(U + K), indicating that turbulence can provide, in principle, ad-
ditional support against gravity. However, note that this assumes that the
turbulence is microscopic (microturbulence), since it is implicitly required
that the velocity field itself does not alter the mass distribution within the
volume.
Magnetic support. Assume now a spherical cloud (a density peak) of ra-
dius R, with a uniform field B = Boex inside and a vanishing field out-
side, and u = P = 0.
It is easy to show that for such a cloud, W =
oR2/6 ≡ φ2/(6π2R), where
−GR [M (r)/r]dM (r) = −3GM 2/5R and M = B2
M (r) is the mass interior to radius r and φ is the magnetic flux, which,
according to the flux-freezing condition, eq. (4), is conserved in the absence
of magnetic dissipation. Since φ is constant, both W and M scale as R−1
upon contraction, and the ratio W/M is constant. This implies that there is
a critical mass-to-flux ratio (M/φ)crit ≡ (5/18)1/2(π2G)−1/2 below which a
cloud cannot collapse gravitationally, since the magnetic energy M is larger
than the gravitational energy W (Strittmatter 1966). Modifications to this
critical ratio due to the geometry of the cloud were studied analytically also
by Strittmatter (1966) and numerically by Mouschovias & Spitzer (1976), in
the presence of an external pressure. Bertoldi & McKee (1992) have studied
the stability of spheroidal, pressure-confined clumps, giving estimates of the
Jeans and magnetic critical masses.
Finally, let us note that throughout this course, we discuss the scalar
version of the VT. A tensor form can also be derived, which allows the treat-
ment of multidirectional transport phenomena. Zweibel (1990) has given a
detailed account of magnetic support and stability of spheroidal, quiescent
cloud configurations using the tensor VT.
3.2. Properties of molecular clouds. Larson's relations
Molecular clouds exhibit a number of correlations between their various phys-
ical properties, such as their size, velocity dispersion, magnetic field strength
and (possibly) density and mass. Larson (1981) first noticed, from a collec-
tion of data from several surveys available to him, that the velocity dispersion
σ and number density n of molecular clouds scaled with their size as
σ
km s−1(cid:17) ∼ 1.1(cid:16) R
pc(cid:17)0.38
,
cm−3(cid:17) ∼ 3400(cid:16) R
(cid:16) n
pc(cid:17)−1.1
.
(12)
(cid:16)
A large number of subsequent studies (e.g., Torrelles et al. 1983; Dame et al.
1986; Falgarone & P´erault 1987; Myers & Goodman 1988a; Falgarone et al.
1992; Miesch & Bally 1994; Wood et al. 1994; Caselli & Myers 1995) have
found similar scalings, although the currently most favored forms are
σ ∼ R1/2
(13)
Virial Theorem in Interstellar Clouds
and
ρ ∼ R−1
9
(14)
where ρ = mn is the mass density, and m is the mean molecular mass. These
scaling relations have been interpreted, since their discovery, as evidence
for virial equilibrium of molecular clouds. Larson (1981) noted that, as a
consequence of these relations, the ratio 2GM/σ2R was roughly constant for
his cloud sample, indicating virial balance between the turbulent velocity
dispersion and self-gravity (W ∼ K). A problem with this result however,
was that the observed linewidths imply in general highly supersonic velocity
dispersions. Therefore, it was expected that such random motions should
produce shocks and dissipate rapidly, becoming subsonic (e.g., Mestel 1965a,
b). This problem disappears, however, if one introduces magnetic fields into
the picture, since Alfv´en waves do not dissipate as rapidly, and additionally
for clouds with nearly the critical mass-to-flux ratio, the Alfv´en velocity
nearly equals the virial velocity (Shu et al. 1987; Mouschovias 1987). That
is, from W ∼ M one gets
B2
4πρ
≡ v2
A ∼
GM
R
∼ GρR2,
while from W ∼ K, we obtain
σ2 ∼ GρR2,
(15)
(16)
where vA is the Alfv´en speed.2 This in principle could solve the dissipation
problem, since in this case velocity dispersions σ up to the Alfv´en speed
do not induce shocks (although see §3.3.), while explaining the observed
virial value of the velocity dispersion. If additionally one assumes that for
an ensemble of clouds the density-size relation ρ ∼ R−1 is satisfied, then
σ2 ∼ GρR3/R ∼ R, automatically satisfying the velocity dispersion-size
relation. That is, out of the two Larson's relations and the virial equilibrium
condition, only two are independent. Furthermore, the relation ρ ∼ R−1 can
be interpreted as a consequence of the clouds being near the critical mass-
to-flux ratio at roughly constant magnetic field strengths. Indeed, from eq.
(15) it follows that ρ ∼ R−1 for constant B.
Myers & Goodman (1988a) tested these ideas on a cloud sample with
existing magnetic field strength determinations from Zeeman measurements.
Assuming σ ∼ vA, it follows from eqs. (15) and (16) that the field strength
necessary for virial equilibrium is Beq ∼ σ2/R. Myers & Goodman (1988a)
plotted the observed magnetic field strength versus Beq, finding the two were
equal to within factors of ∼ 3 (fig. 2).
2Xie (1997) has pointed out that it is actually the magnetic fluctuations ∆B that should be
considered in these expressions, rather than the mean field B. However, the two generally
10
Enrique V´azquez-Semadeni
Figure 1. Observed vs. virial equilibrium values of the magnetic
field for the cloud sample of Myers & Goodman 1988a.
Virial Theorem in Interstellar Clouds
11
Figure 2. Observed magnetic field strength vs. virial density
n ≡ 15(32π ln 2mG)−1(∆v/R)2 for the cloud sample of Myers &
Goodman 1988a. The solid lines indicate the virial equilibrium
model for the indicated velocity dispersions.
Additionally, they also found that Bobs ∼ ρ1/2, in agreement with σ2 ∼
B2/ρ (again derived from combining relations [15] and [16]), since in their
sample σ had relatively little scatter compared with the other quantities.
The resulting correlation had a scatter consistent with the scatter in σ (fig.
2). Note, however, that in the latter plot, the "virial" density they used
was an estimate assuming virial equilibrium (ρ ∝ (∆v/R)2), not an actual
observed value, and thus this plot can only be taken as a proof of self-
consistency of the virial equilibrium hypothesis.
Myers & Goodman (1988b) also considered the fact that the observed
linewidths contain both a thermal and a nonthermal component, explaining
deviations from Larson's (1981) trends at very small scales (cloud cores of
sizes ∼ 0.1 pc). According to relation (13), which refers to the macroscopic,
nonthermal velocity dispersion, at small enough cloud sizes the nonthermal
component becomes smaller than the thermal (the velocity dispersion be-
comes subsonic).
Indeed, Myers & Goodman (1988b) find a drop in the
ratio of nonthermal to gravitational energy (KNT/W ) in their cloud sample
for small cloud sizes.
It is important to note, however, that the trend is
marginal when the raw data are used, but clearly noticeable only when the
assumption of virial equilibrium is introduced (replacing the true gravita-
tional energy obtained from direct measurements of the size and density by
the total kinetic (thermal + nonthermal) energy). From this result, Myers
turn out to be of comparable magnitude (Myers & Goodman 1991), and use of B is
justified, alhough the difference should be kept in mind.
12
Enrique V´azquez-Semadeni
& Goodman (1988b) suggest that the uncertainties in KNT/W arise more
from uncertainties in the density and size than from uncertainties in velocity
dispersion and tempertaure. However, an obvious alternative possibility is
that the scatter is real, indicating departures from precise virial equilibrium.
Their plot "assuming virial equilibrium" is in reality just a plot of the ratio
of nonthermal to total kinetic energy, thus necessarily being very tight at
large scales if the velocity dispersion does increase with size while the tem-
perature remains roughly constant, but it is not a test of virial equilibrium
in the clouds. On the other hand, further supporting evidence for virial
equilibrium has been provided by Goodman et al. (1993), who find good
agreement between the observed and the "virial" values of β, the ratio of
rotational to gravitational energy in a sample of dense cores.
The role of thermal plus nonthermal (TNT) motions has been further
investigated by Fuller & Myers (1992) and by Caselli & Myers (1995), finding
respectively that the scaling of the nonthermal part of the velocity differs
from that of the total velocity dispersion, and that the scaling is different for
massive cores than for low-mass cores. Furthermore, Fuller & Myers (1992)
have distinguished between cloud-to-cloud scaling, i.e., that obtained when
the sample consists of various clouds observed with the same tracer (sensitive
to one particular density range) and line-to-line scaling, i.e., the scaling
obtained when using various tracer molecules (sensitive to a variety of density
ranges) on a single cloud. The latter can be used to test for virialization
and the density structure of individual clouds as a function of the distance
from the core. An even finer subdivision of possibilities has been proposed
by Goodman et al.
(1997). These authors have additionally found that
the slope of the σ-R relation using single-line, single-cloud measurements,
appears to decrease to nearly zero at scales ∼ 0.1 pc, and that the column
density N changes its size dependence from N ∼ R−0.2 at scales >∼ 0.1 pc to
N ∼ R−0.9 at scales <∼ 0.1 pc. They interpret these results as evidence for
a transition from non-thermal to predominantly thermal support, and the
onset of "velocity coherence" at those scales.
Finally, it should be noted that a number of other possible mechanisms,
capable of originating Larson's relations or either one thereof, have been
proposed. For example, Larson (1981) himself suggested that the density-
size relation might be due to planar shocks which keep the column density
constant along the direction of shock propagation. Chi`eze (1987) has pro-
posed that an ensemble of clouds on the verge of gravitational instability
in an environment providing external pressure should also satisfy Larson's
relations. In the chapter on Turbulence in Molecular Clouds, a discussion on
mechanisms related to hydro- or magnetohydrodynamic turbulence is given.
Other scenarios are unfortunately out of the scope of this course.
Virial Theorem in Interstellar Clouds
13
Nonthermal
Figure 3.
Myers & Goodman 1988a. No clear correlation is seen.
Mouschovias & Psaltis 1995.)
linewidth vs. size for 14 objects from
(From
3.3. Counterpoint
As it may have started to be apparent to the reader from the discussion of the
previous section, the issue of cloud virialization as the origin of the Larson
(1981) relations is not completely settled. Besides the caveats inherent to the
work of Myers & Goodman (1988a, b), which are the standard references for
observational determinations of virial equilibrium in clouds, other somewhat
contradictory evidence exists.
First, let us distinguish between a cloud being in virial equilibrium and
it satisfying the VT. Actually, all clouds, and in general all fluid parcels
in the ISM, satisfy the VT, as it is a direct consequence of the momen-
tum conservation equation. Instead, virial equilibrium means d2I/dt2 = 0
specifically.
Concerning the velocity dispersion-size relation, Scalo (1990) and Mous-
chovias & Psaltis (1995) have pointed out that the original compiled by
Myers & Goodman (1988a) show essentially just scatter for this relation (fig.
3.3.). Mouschovias & Psaltis (1995) "solve the problem" by plotting instead
the ratio σ/R1/2 vs. the magnetic field strength B. In reality, however, this is
nothing but the same Bobs vs. Beq plot shown by Myers & Goodman (1988a).
Moreover, Mouschovias & Psaltis (1995) use this result as an argument in
favor of the σ-R relation being a consequence of magnetic support. However,
they arrive at this conclusion by assuming B does not "vary much from
place to place in the ISM under conditions suitable for the formation of
self-gravitating clouds"; yet, the magnetic field strength data they consider
vary by three orders of magnitude! In summary, the cloud data discussed
in Myers & Goodman (1988a) are consistent with magnetic support, but do
not satisfy the dispersion-size relation.
14
Enrique V´azquez-Semadeni
Furthermore, there exists both observational and numerical evidence
that the density-size relation is not always satisfied, implying that the col-
umn density is not constant. It was already pointed out by Larson (1981)
himself that the density-size relation might not be a real property of molec-
ular clouds, but instead just an artifact of limitations of the observational
methods used to produce cloud surveys. Kegel (1989) and Scalo (1990)
elaborated on this issue. Also, a number of observational studies have found
significantly different scaling exponents for both Larson's relations, or no
clear scaling at all. For example, Carr (1987) has found σ ∼ R0.25 and
M ∼ R2.5, implying ρ ∼ R−0.5 for 45 clumps in the Cep OB 3 molecular
cloud. In fact, Carr notes that the clumps are not in virial equilibrium, and
concludes that they must be expanding. Loren (1989) studied 89 clumps in
the ρ Oph molecular cloud, finding ρ ∼ R−0.2 and no σ-R correlation. He
points out that many of the clumps are extremely filamentary and that they
may probably be the outcome of the passage of a shock wave. Falgarone et
al. (1992) focused on low brightness areas of several molecular clouds, includ-
ing both gravitationally bound and unbound structures (structures for which
the actual masses are much smaller than the value obtained assuming virial
equilibrium), finding σ ∼ R0.4, and a density-size plot with large scatter and
an upper bound given by a Larson-type relation ρ ∼ R−1. V´azquez-Semadeni
et al. (1997) have found the same effect in numerical simulations of turbu-
lent cloud formation in the ISM, in which in fact there is no unique ρ-R
correlation. Plume et al. (1997), in a sample of maser-selected, massive star
formation regions, find no significant σ-R or ρ-R correlations either.
Finally, Myers et al. (1995) have found an example of a cloud in which
the kinetic and magnetic energies are comparable, yet the gravitational en-
ergy is roughly 500 times smaller, indicating that self-gravity is negligible
for this cloud.
4. Discussion and conclusions
From the discussion in the previous sections, a number of points can be
concluded.
1. "Virial equilibrium", i.e., d2I/dt2 = 0 allows up to a constant rate of
change in the moment of inertia I, even though it is commonly taken to
mean a stationary equilibrium.
2. Larson's (1981) relations, eqs. (13) and (14), have been interpreted as
a consequence of virial equilibrium3 between self-gravity on the one hand
and turbulent velocity dispersion and/or magnetic support on the other.
3Mouschovias & Psaltis (1995) distinguish between magnetic virial equilibrium and mag-
netic support, but the resulting balance equation is the same.
Virial Theorem in Interstellar Clouds
15
From the observational data discussed above, this interpretation may apply
to strongly condensed clumps which, due to self-gravity, have decoupled
from their surrounding medium, and thus require internal support against
collapse.
3. The above interpretation, however, is insufficient for explaining both scal-
ing relations. One is still independent and requires an additional explanation.
A critical mass-to-flux ratio has been invoked as the origin of the density-size
relation assuming roughly constant magnetic field strengths (Shu et al. 1987;
Myers & Goodman 1988b; Mouschovias & Psaltis 1995), but this assump-
tion is not verified in the observational data. For example, the magnetic
field strenghts discussed by Myers & Goodman (1988a) and Mouschovias
and Pslatis (1995) span three orders of magnitude. Besides, examples of
virialized clouds which however follow different scaling relations exist (e.g.,
Loren 1989; Fuller & Myers 1992; Caselli & Myers 1995; Goodman et al.
1997; see also V´azquez-Semadeni & Gazol 1995 for a theoretical discussion).
Planar shocks have also been invoked as an explanation of the apparently
constant column densities (Larson 1981; Scalo 1987).
4. It is possible that the density-size relation is not a true property of clouds,
but only an observational effect due to the limited integration times used in
surveys (Larson 1981; Kegel 1989; Scalo 1990), which tend to select constant-
column density obejcts. In this case, surveys using larger integration times
would exhibit an increasingly larger scatter in density-size plots, as in the
data of Falgarone et al. (1992) or the numerical data of V´azquez-Semadeni
et al. (1997), the latter being free from such limitations. In fact, large scat-
ter was already present in the Myers & Goodman (1988b) data. Although
they interpreted it as uncertainty in the data rather than real scatter about
virialization, the latter possibility is equally feasible. A study that seemed to
find exceptionally constant column densities while claiming a large dynamic
range (Wood et al. 1994) has been questioned by V´azquez-Semadeni et al.
(1997).
5. The velocity dispersion-size relation does not appear subject to the pos-
sibility of a spurious origin, and therefore it is probably real when detected,
although in many cases it is not (e.g., Loren 1989; Plume et al. 1997). For
the cases when it is present, a number of physical mechanisms are plausible
candidates. If the density-size relation is not verified, then the standard ar-
guments based on virial equilibrium between gravity and velocity dispersion
cannot be invoked (recall the σ-R relation is a consequence of virial equi-
librium and the density-size relation). In particular, the possibility that the
σ-R relation is a consequence of the characteristic energy spectrum of an
ensemble of shocks in the flow has been suggested by a number of authors
(e.g., Passot et al. 1988; Padoan 1995; Gammie & Ostriker 1996; Fleck 1996;
V´azquez-Semadeni et al. 1997). In this case, the origin of the dispersion-size
16
Enrique V´azquez-Semadeni
relation would be completely independent of the virial condition, explaining
why the relation can be present even in the absence of a density-size relation.
6. In summary, the Larson's relations and their virial equilibrium interpre-
tation probably apply to relaxed, strongly self-gravitating clumps. Regions
which do not exhibit clear scaling relations often include either massive cores
(Caselli & Myers 1995; Plume et al. 1997) or regions with strong evidence
of recent perturbations (Loren 1989).
In fact, the dense cores studied by
Plume et al. were selected by the presence of H2O masers, suggesting strong
excitation mechanisms as well. Such "perturbed" regions are likely not in
virial equilibrium, thus also being transients (e.g., Magnani et al. 1993), or
at least having strongly fluctuating moments of inertia (shapes and internal
mass distributions).
Acknowledgments. The author wishes to thank A. Goodman, T. Pas-
sot and J. Scalo for a critical reading of the manuscript, as well as helpful
comments and precisions.
References
Ballesteros-Paredes & V´azquez-Semadeni 1997 in preparation
Bally, J., Langer, W. D., Stark, A. A. & Wilson, R. W. 1987, ApJ, 312, L45
Bertoldi, F., & McKee, C. F. 1992, ApJ, 395, 140
Bonazzola, S., Falgarone, E., Heyvaerts, J., P´erault, M., Puget, J. L. 1987,
A&A 172, 293
Carr, J. S. 1987, ApJ, 323, 170
Caselli, P. & Myers, P. C. 1995, ApJ, 446, 665
Chandrasekhar, S. 1951, Proc. R. Soc. London, 210, 26
Chi`eze, J. P. 1987, A& A, 171, 225
Cowling, T. G. 1976, Magnetohydrodynamics (Bristol: Adam Hilger Ltd.)
Currie, I. G., 1974, Fundamental Mechanics of Fluids (New York: McGraw-
Hill)
Dame, T., Elmegreen, B. G., Cohen, R., Thaddeus, P. 1986, ApJ, 305, 892
Falgarone, E., & P´erault, M. 1987,
in Physical Processes in Interstellar
Clouds, ed. G. Morfill, & M. Scholer (dordrecht: Reidel), 59
Falgarone, E., Phillips, T. G., & Walker, C. K. 1991, ApJ, 378, 186
Falgarone, E., Puget, J.-L., & P´erault, M. 1992, A&A, 257, 715
Fleck, R. C. 1996, ApJ, 458, 739
Fuller, G. A., & Myers, P. C. 1992, ApJ, 384, 523
Gammie, C. F. & Ostriker, E. 1996, ApJ, 466, 814
Virial Theorem in Interstellar Clouds
17
Goodman, A. A., Benson, P. J., Fuller, G. A. & Myers, P. C. 1993, ApJ,
406, 528
Goodman, A. A., Barranco, J. A., Wilner, D. J. & Heyer, M. H. 1997, ApJ,
submitted
Jura, M. 1987,
in Interstellar Processes, ed. D. J. Hollenbach & H. A.
Thronson (Dordrecht: Reidel), 3
Kegel, W. H. 1989, A&A, 225, 517
Landau, L. D. & Lifshitz, E. M. 1987, Fluid Mechanics, 2nd Ed. (Oxford:
Pergamon Press)
Larson, R. B. 1981, MNRAS, 194, 809
L´eorat, J., Passot, T., & Pouquet, A. 1990, MNRAS, 243, 293
Lesieur, M. 1990, Turbulence in Fluids, 2nd ed. (Dordrecht: Kluwer)
Loren, R. B. 1989, ApJ, 338, 902
Magnani, L., LaRosa, T. N., & Shore, S. N. 1993, ApJ, 402, 226
McKee, C. F., Zweibel, E. G. 1992, ApJ, 399, 551 (MZ92)
Mestel, L., & Spitzer, L. 1956, MNRAS, 116, 503
Mestel, L. 1965a, QJRAS, 6, 161
Mestel, L. 1965b, QJRAS, 6, 265
Miesch, M. S., & Bally, J. 1994, ApJ, 429, 645
Mouschovias, T. Ch. 1987, in Physical Processes in Interstellar Clouds, ed.
G. E. Morfill & M. Scholer (Dordrecht: Reidel)
Mouschovias, T. Ch. & Spitzer, L., Jr. 1976, ApJ, 210 326
Mouschovias, T. Ch. & Psaltis, D. 1995, ApJ, 444, L105
Myers, P. C., & Goodman, A. A. 1988a, ApJ, 326, L27
Myers, P. C., & Goodman, A. A. 1988b, ApJ, 329, 392
Myers, P. C., & Goodman, A. A. 1991, ApJ, 373, 509
Myers, P. C., Goodman, A. A., Gusten, R. & Heiles, C. 1995, ApJ, 442, 177
Padoan, P. 1995, MNRAS, 277, 377
Passot, T. V´azquez-Semadeni, E., & Pouquet, A. 1995, ApJ, 455, 536
Passot, T., Pouquet, A. & Woodward P., 1988, A&A, 197, 228
Plume, R., Jaffe, D. T., Evans, N. J. II, Mart´ın Pintado, J., & G´omez-
Gonz´alez, J. 1997, ApJ, in press
Scalo, J. M. 1987, in Interstellar Processes, ed. D. J. Hollenbach and H. A.
Thronson (Dordrecht: Reidel), 349
Scalo, J. M. 1990, in Physical Processes in Fragmentation and Star Forma-
tion, ed. R. Capuzzo-Dolcetta, C. Chiosi, & A. di Fazio (Dordrecht:
Kluwer), 151
Shore, S. N. 1992, An Introduction to Astrophysical Hydrodynamics (San
Diego: Academic Press)
18
Enrique V´azquez-Semadeni
Shu, F. 1992, Gas Dynamics (Mill Valley: University Science Books)
Shu, F., Adams, F. C., Lizano, S. 1987, ARAA, 25, 23
Spitzer, L., Jr., 1978, Physical Processes in the Interstellar Medium (New
York: Wiley-Interscience)
Strittmatter, P. A. 1966, MNRAS, 132, 359
Torrelles, J. M., Rodr´ıguez, L. F., Cant´o, J., Carral, P., Marcaide, J., Moran,
J. M., & Ho, P. T. P. 1983, ApJ, 274, 214
V´azquez-Semadeni, E., Ballesteros-Paredes, J. & Rodr´ıguez, L. F. 1997, ApJ,
474, 292
V´azquez-Semadeni, E. & Gazol, A. 1995, A&A, 303, 204
V´azquez-Semadeni, E., Passot, T. & Pouquet, A. 1995, ApJ, 441, 702
Wood, D. O. S., Myers, P. C., & Daugherty, D. 1994, ApJS, 95, 457
Xie, T. 1997, ApJ, in press
Zweibel, E. G. 1990, 348, 186
|
astro-ph/9809335 | 2 | 9809 | 1999-02-09T16:09:14 | The Disk-Magnetosphere Interaction in the Accretion-Powered Millisecond Pulsar SAX J1808.4-3658 | [
"astro-ph"
] | The recent discovery of the first known accretion-powered millisecond pulsar with the Rossi X-Ray Timing Explorer provides the first direct probe of the interaction of an accretion disk with the magnetic field of a weakly magnetic (B<10^10 G) neutron star. We demonstrate that the presence of coherent pulsations from a weakly magnetic neutron star over a wide range of accretion rates places strong constraints on models of the disk-magnetosphere interaction. We argue that the simple Mdot^(3/7) scaling law for the Keplerian frequency at the magnetic interaction radius, widely used to model disk accretion onto magnetic stars, is not consistent with observations of SAX J1808.4-3658 for most proposed equations of state for stable neutron stars. We show that the usually neglected effects of multipole magnetic moments, radiation drag forces, and general relativity must be considered when modeling such weakly magnetic systems. Using only very general assumptions, we obtain a robust estimate of mu~(1-10)x10^26 G cm^3 for the dipole magnetic moment of SAX J1808.4-3658, implying a surface dipole field of ~10^8-10^9 G at the stellar equator. We therefore infer that after the end of its accretion phase, this source will become a normal millisecond radio pulsar. Finally, we compare the physical properties of this pulsar to those of the non-pulsing, weakly magnetic neutron stars in low-mass X-ray binaries and argue that the absence of coherent pulsations from the latter does not necessarily imply that these neutron stars have significantly different magnetic field strengths from SAX J1808.4-3658. | astro-ph | astro-ph |
The Disk-Magnetosphere Interaction in the Accretion-Powered Millisecond
Pulsar SAX J1808.4 -- 3658
Dimitrios Psaltis
Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138; [email protected]
and
Deepto Chakrabarty
Department of Physics and Center for Space Research, Massachusetts Institute of Technology,
Cambridge, MA 02139; [email protected]
To appear in The Astrophysical Journal.
ABSTRACT
The recent discovery of the first known accretion-powered millisecond pulsar with
the Rossi X-Ray Timing Explorer provides the first direct probe of the interaction of
an accretion disk with the magnetic field of a weakly magnetic (B ∼< 1010 G) neutron
star. We demonstrate that the presence of coherent pulsations from a weakly magnetic
neutron star over a wide range of accretion rates places strong constraints on models
of the disk-magnetosphere interaction. We argue that the simple M 3/7 scaling law for
the Keplerian frequency at the magnetic interaction radius, widely used to model disk
accretion onto magnetic stars, is not consistent with observations of SAX J1808.4 -- 3658
for most proposed equations of state for stable neutron stars. We show that the usually
neglected effects of multipole magnetic moments, radiation drag forces, and general
relativity must be considered when modeling such weakly magnetic systems.
Using only very general assumptions, we obtain a robust estimate of µ ≃ (1 − 10) ×
1026 G cm3 for the dipole magnetic moment of SAX J1808.4 -- 3658, implying a surface
dipole field of ∼ 108 − 109 G at the stellar equator. We therefore infer that after the
end of its accretion phase, this source will become a normal millisecond radio pulsar.
Finally, we compare the physical properties of this pulsar to those of the non-pulsing,
weakly magnetic neutron stars in low-mass X-ray binaries and argue that the absence of
coherent pulsations from the latter does not necessarily imply that these neutron stars
have significantly different magnetic field strengths from SAX J1808.4 -- 3658.
Subject headings: accretion, accretion disks -- pulsars: individual: SAX J1808.4 -- 3658
-- stars: neutron -- X-rays: stars
-- 2 --
1.
INTRODUCTION
The basic framework for understanding accretion-powered pulsars emerged soon after their
discovery (Giacconi et al. 1971). These systems are rotating neutron stars accreting matter from
a binary companion, with magnetic fields strong enough to disrupt the accretion flow above the
stellar surface (Pringle & Rees 1972; Davidson & Ostriker 1973; Lamb, Pethick, & Pines 1973).
When threaded by the stellar magnetic field, the accreting gas is brought into corotation with the
star and is channeled along field lines to the polar caps, releasing its potential and kinetic energy
mostly in X-rays. The rotation of these hot spots through our line of sight produces X-ray pulses
at the spin frequency of the neutron star. Most of the ≈ 50 accretion-powered pulsars have spin
frequencies and inferred dipole magnetic field strengths in the range ∼1 − 103 s and ∼1011 − 1013 G,
respectively (White, Nagase, & Parmar 1995).
In strongly magnetic (B ∼> 1011 G) accreting neutron stars, models for the disk-magnetosphere
interaction based on this framework (see Ghosh & Lamb 1991 for a review) are generally consistent
with the accretion torque behavior of the Be/X-ray pulsar transients, as observed by the Comp-
ton/BATSE all-sky monitor (see Bildsten et al. 1997 and references therein). The Be/X-ray pulsar
transients allow for a direct test of the predicted scaling of torque with accretion rate, since they
span a wide range of accretion rates during their outbursts. However, the same models can only ac-
count for the bimodal torque reversals observed in several persistent accreting pulsars if additional
assumptions are introduced (Chakrabarty et al. 1997a, 1997b; Bildsten et al. 1997). Examples
of suitable assumptions are a bimodal distribution of the mass transfer rate onto the pulsar, or
a bimodal dependence on accretion rate of the orientation of the disk (Nelson et al. 1997; van
Kerkwijk et al. 1998), of the orbital angular velocity of the accreting gas (Yi & Wheeler 1998), or
even of the strength and orientation of any magnetic field produced in the disk (Torkelsson 1998).
In weakly-magnetic (B ∼< 1010 G) accreting neutron stars, models of the disk-magnetosphere
interaction have only been tested indirectly so far. Most neutron stars in low-mass X-ray binaries
(LMXBs) show no periodic oscillations in their persistent emission. They have dipole magnetic fields
∼< 1010 G, as inferred from their bursting (Lewin, van Paradijs, & Taam 1995) and rapid variability
behavior (see van der Klis 1998 for a review) as well as from their X-ray spectra (Psaltis & Lamb
1998). Their power-density spectra show various types of quasi-periodic oscillations (QPOs), and in
particular the most luminous of the neutron-star LMXBs show ∼ 15−60 Hz QPOs. These are called
horizontal-branch oscillations (HBOs) and have frequencies that increase with mass accretion rate
(van der Klis et al. 1985; van der Klis 1989). They have been interpreted as occurring at the beat
frequency between the Keplerian frequency at the magnetic interaction radius (where the stellar
magnetic field strongly interacts with the accretion disk) and the neutron star spin frequency (Alpar
& Shaham 1985; Lamb et al. 1985). Models of the inner accretion disk and the disk-magnetosphere
interaction can account for the observed scaling of HBO frequency with accretion rate if the neutron
stars in these systems are near their magnetic spin equilibrium and if the inner accretion disks are
radiation-pressure dominated (Psaltis et al. 1999b).
-- 3 --
The discovery of ∼ 200 − 1200 Hz QPOs (often occurring in pairs; hereafter kHz QPOs; see
van der Klis 1998 and references therein) in the X-ray flux of many neutron-star LMXBs and the
identification of the higher-frequency QPO with a Keplerian orbital frequency in the accretion disk
(van der Klis et al. 1996; Strohmayer et al. 1996; Miller, Lamb, & Psaltis 1998) has introduced
additional complications in studying the disk-magnetosphere interaction using HBO observations
(see also Psaltis, Belloni, & van der Klis 1999a). The magnetic interaction radius inferred from
the HBO frequencies is larger than the disk radius responsible for the higher-frequency kHz QPO.
Therefore, if the higher-frequency kHz QPO is a Keplerian orbital frequency in the disk, then the
magnetospheric beat-frequency model for the HBO can be valid only if a non-negligible fraction
of the disk plasma is not threaded by the stellar magnetic field but remains in the disk plane
inside the interaction radius (see Miller et al. [1998] and Psaltis et al. [1999b] for a discussion).
This requirement has not yet been addressed in any theoretical model for the disk-magnetosphere
interaction and therefore cannot be tested directly.
The recent discovery of the first weakly-magnetic accretion-powered pulsar with the Rossi X-
ray Timing Explorer (RXTE) allows the first direct test of disk-magnetosphere interaction models
in the weak-field limit. SAX J1808.4 -- 3658 is a transient X-ray source that shows type I X-ray
bursts (in't Zand et al. 1998) and coherent 401 Hz X-ray pulsations (Wijnands & van der Klis
1998a), and is a member of a low-mass binary in a 2-hour orbit (Chakrabarty & Morgan 1998).
Its distance is ≈4 kpc, as inferred from flat-topped type I X-ray bursts that are thought to be
Eddington limited (in 't Zand et al. 1998), and its luminosity varied by ∼>2 orders of magnitude
during the 1998 April/May outburst (Cui, Morgan, & Titarchuk 1998).
In this paper we study the disk-magnetosphere interaction in SAX J1808.4 -- 3658. In §2, we com-
pare the predictions of theoretical models for the inner accretion disk and the disk-magnetosphere
interaction with observations of SAX J1808.4 -- 3658 and, in §3, we infer the magnetic field strength
of the pulsar. In §4, we discuss our results and their implications for disk-magnetosphere interac-
tion models and compare SAX J1808.4 -- 3658 to the millisecond radio pulsars and the non-pulsing
neutron stars in LMXBs.
2. DISK-MAGNETOSPHERE INTERACTION
2.1. Assumptions and Formalism
Throughout this paper we assume that most of the accreting gas around SAX J1808.4 -- 3658
is confined in a geometrically thin accretion disk before interacting with the pulsar magnetic field.
This assumption is supported by the transient nature of the source, which can be understood in
terms of dwarf-nova -- like accretion disk instabilities (see Chakrabarty & Morgan 1998). We also
neglect the effect of wind mass loss from the inner accretion disk and of radiation drag forces, as
well as all general relativistic effects.
-- 4 --
The radius r0 at which magnetic stresses remove the angular momentum of the disk flow and
disrupt it can be estimated by balancing the magnetic and material stresses, (Ghosh & Lamb 1978,
1979a)
BpBφ
4π
4πr2
0∆r0 = M Ωr2
0 ,
(1)
where Bp and Bφ are the poloidal and toroidal components of the magnetic field, ∆r0 is the radial
M is the mass transfer rate through the inner disk, and Ω(r) is the
width of the interaction region,
angular velocity of the gas at radius r. Assuming that the poloidal magnetic field is dipolar with
magnetic moment µ and that the accretion flow is Keplerian, we obtain
r0 = γ2/7
B µ4
GM M 2!1/7
,
(2)
where γB ≡ (Bφ/Bp)(∆r0/r0), M is the mass of the neutron star, and G is the gravitational
constant.
The boundary layer parameter γB in equation (2) depends in general on all the other phys-
ical quantities. Hence, different models for the inner accretion disk's environment and the disk-
magnetosphere interaction generally predict different coefficients and scalings (Ghosh & Lamb 1992;
see also Ghosh & Lamb 1978, 1979a; Wang 1996; Ghosh 1996). However, if the disk-magnetosphere
interaction takes place in a region of the accretion disk where all physical quantities have a power-
law dependence on radius (as is the case for most accretion disk models away from the stellar
surface) and equation (1) describes angular momentum balance in the interaction region, then the
Keplerian orbital frequency at r0 can be written as
ν0 ≃ νK,0(cid:18) M
M⊙(cid:19)γ(cid:18)
µ
1027 G cm3(cid:19)β M
ME!α
,
(3)
ME = 2 × 10−8 M⊙ yr−1 is the Eddington critical accretion rate (at
where M⊙ is the solar mass;
which the outward radiation forces in a spherically symmetric hydrogen flow balance gravity); and
α, β, γ, and νK,0 are parameters which depend upon the nature of the inner accretion disk. For
our study, we used the parameters from four different inner disk models (see Table 1). We chose
these models both because they span a wide range of physical conditions in the inner disk (one-
and two-temperature plasmas, and gas and radiation-pressure -- dominated flows) and a range of
radiation processes, and because detailed calculations of the interaction with a stellar magnetic
field have been performed for these models (see Table 1 and Ghosh & Lamb 1992 for a discussion
of the physical processes involved in these models).
2.2. Limits from the Detection of Coherent Pulsations
We can use the presence of coherent X-ray pulsations in SAX J1808.4 -- 3658 over a wide range
of mass accretion rate to test scaling (3) in the weak magnetic field regime. For a rotating neutron
-- 5 --
star to appear as an accretion-powered pulsar, the stellar magnetic field must be strong enough to
disrupt the Keplerian disk flow above the stellar surface1. Note that the orbital frequency ν0 at the
interaction radius r0 increases with mass accretion rate (α > 0). Therefore, the stellar magnetic
field must be strong enough to disrupt the disk flow at radii larger than the neutron star radius
R at the maximum mass accretion rate Mmax for which coherent pulsations were detected. This
leads to a lower limit on the magnetic dipole moment,
µ ≥ 1027 4π2ν2
K,0R3
GM⊙ !−1/2β
M⊙(cid:19)(1−2γ)/2β Mmax
(cid:18) M
ME !−α/β
G cm3 .
(4)
At the same time, the stellar magnetic field must be weak enough that accretion is not centrifugally
inhibited at the minimum mass accretion rate Mmin for which coherent pulsations were detected.
Therefore, the orbital frequency ν0 at this accretion rate must be larger than the spin-frequency of
the neutron star νs. This leads to an upper limit on the magnetic dipole moment,
µ ≤ 1027 νs
νK,0!1/β
M⊙(cid:19)−γ/β Mmin
(cid:18) M
ME !−α/β
G cm3 .
(5)
Of course, X-ray brightness modulations at the pulsar spin frequency may be possible even if
one of the above two requirements is not met. For example, azimuthal variations in the efficiency
of processes dependent on magnetic field strength (such as cyclotron emission, absorption, and
resonant scattering) may be strong enough to produce X-ray pulsations, even if the magnetic field
of the neutron star is dynamically unimportant and the accretion disk extends down to the stellar
surface. However, if at any point during the outburst either the disk had reached the stellar
surface or accretion had been centrifugally inhibited, then either the X-ray spectrum or the pulse
amplitude should have changed abruptly. Instead, the X-ray spectrum of SAX J1808.4 -- 3658 was
found to be remarkably stable (Gilfanov et al. 1998; Heindl & Smith 1998), and the pulse amplitude
to increase only slightly from 4% to 7% as the inferred luminosity declined by more than two orders
of magnitude during the outburst (Cui et al. 1998).
Combining constraints (4) and (5) we obtain
M
M⊙
> 0.047(cid:18) νs
401 Hz(cid:19)2 Mmax
Mmin!2α
10 km(cid:19)3
(cid:18) R
.
(6)
Given the observed spin frequency and range of accretion rates at which coherent pulsations are
detected, inequality (6) defines a limiting curve in the M -R parameter space for neutron stars. The
scaling of this curve is cubic in R, independent of the details of the disk-magnetosphere interaction
1Note that this is a necessary but not a sufficient condition: the X-ray brightness at infinity may not be modulated
with any detectable amplitude if, e.g., the spin of the neutron star is perfectly aligned to its magnetic moment, or the
neutron star is surrounded by a scattering medium that attenuates the pulsations (Brainerd & Lamb 1986; Kylafis
& Phinney 1989).
-- 6 --
Fig.
1. -- Constraints on the
neutron star mass and radius in
SAX J1808.4 -- 3658. The solid lines
represent mass-radius relations of
stable neutron stars for various pro-
posed equations of state (Cook et
al. 1994). The dashed lines repre-
sent the limits imposed by various
models for the inner accretion disk
(Table 1). The allowed region for
each inner disk model is to the left
of the corresponding dashed curve.
The limiting curve for the 2S model
essentially lies on the R = 0 axis on
this scale, and is not plotted.
model, with a coefficient that depends on a single model parameter2. As mentioned above, the
X-ray spectrum of SAX J1808.4 -- 3658 was found to be remarkably stable during the outburst, so
we can assume here that the ratio Mmax/ Mmin scales as the ratio of the corresponding countrates in
a given X-ray bandpass, independent of (unknown) bolometric corrections. For the 1998 April/May
outburst of SAX J1808.4 -- 3658, this ratio was ≈ 130 (Cui et al. 1998).
Figure 1 compares the M -R relations for several proposed equations of state for stable neutron
stars (see Cook, Shapiro, & Teukolsky 1994) to the limit (6) for the four previously mentioned inner
disk models applied to SAX J1808.4 -- 3658. Only the limiting curve for the 1R model, for which
α ≃ 0.2, is consistent with many proposed equations of state. In a thin accretion disk, radiation
pressure dominates over gas pressure at radii (see Treves, Maraschi, & Abramowicz 1988)
r
R
≪ 0.5α0.1(cid:18) η
500(cid:19)(cid:18) M
M⊙(cid:19)0.42(cid:18) R
10 km(cid:19)−0.24
M
10−4
ME!0.76
,
(7)
where η is a parameter in the range 102 -- 103 and α ∼< 1 is the Shakura-Sunyaev viscosity parameter.
During the decline phase of the outburst, the mass accretion rate was so small that r/R ≪ 1
and the inner accretion disk in SAX J1808.4 -- 3658 was certainly not radiation-pressure dominated.
2While this work was in progress, we learned of a paper by Burderi & King (1998) in which a similar argument
is used to constrain R by assuming a particular disk-magnetosphere interaction model (essentially the 1G model)
and hence a specific value for α. However, the applicability of the 1G model to SAX J1808.4 -- 3658 is dubious, as we
demonstrate here. In contrast, we use inequality (6) to constrain disk-magnetosphere interaction models and hence
α.
-- 7 --
Whether or not the disk was radiation-pressure dominated at the peak of the outburst is not clear
given the uncertainty in the model parameter η. However, it is very unlikely that there was a
transition between a radiation-pressure and a gas-pressure -- dominated inner accretion disk during
the outburst, given the stability of the X-ray spectrum of the source and of the pulsed fraction of
the emission (see discussion below). Radiation-pressure -- dominated models for the inner accretion
disk, like the one shown in Figure 1, are therefore not applicable to SAX J1808.4 -- 3658 since its
inferred mass accretion rate is significantly sub-Eddington (in't Zand et al. 1998).
From the three gas-pressure dominated models, the two-temperature ones (2B and 2S) are
inconsistent with any of the proposed equations of state and are thus ruled out for SAX J1808.4 --
3658. The one-temperature gas-pressure dominated model (1G) is barely consistent with a very
restricted range of M and R allowed by only a few equations of state. However, the maximum
and minimum accretion rates at which RXTE detected coherent pulsations from SAX J1808.4 --
3658 were determined by the peak luminosity of the outburst and the instrumental sensitivity,
respectively, which are obviously unrelated to the limiting equalities in (4) and (5). This makes it
quite unlikely that the actual mass and radius of SAX J1808.4 -- 3658 lies very close to the limiting
curve 1G in Figure 1. Therefore, either a remarkable coincidence has occurred or (more probably)
the 1G model is also ruled out for SAX J1808.4 -- 3658. This is particularly interesting because the
dependence of ν0 on M predicted by the 1G model is very similar to the simple ν0 ∝ M 3/7 scaling
law derived from equation (2) for constant γB. This simple scaling law is widely used to describe
the disk-magnetosphere interaction in accreting magnetic stars (and is in fact generally considered
the standard model). Figure 1 implies that the scaling of ν0 with M for SAX J1808.4 -- 3658 should
be weaker than predicted by this simple scaling law. In particular, we find that α must be < 0.4,
and hence γB must be a function of
M .
2.3. Predicted Limits on the Accretion Torque in SAX J1808.4 -- 3658
An independent test of scaling (3) in the weak magnetic field regime is possible if the spin
frequency derivative νs in SAX J1808.4 -- 3658 due to accretion torques can be measured. For a
Keplerian disk flow terminated at a radius r0, the accretion torque on the neutron star is given by
2πI νs = η M (GM r0)1/2 ,
(8)
where I is the neutron star moment of inertia and η is a dimensionless quantity in which all
the physics of the disk-magnetosphere interaction is parametrized (Ghosh & Lamb 1979b). For a
prograde accretion disk, η is positive and a strong function of both the magnetic field strength and
the mass accretion rate (Ghosh & Lamb 1979b). For a retrograde accretion disk η is negative and
has an absolute value of order unity (Daumerie 1996).
Applying constraints (4) and (5) to equation (8), we obtain an upper and a lower limit on the
spin frequency derivative for SAX J1808.4 -- 3658 predicted by accretion torque theory,
-- 8 --
νs ≤ 1.8 × 10−13η I/1045 g cm2
M/M⊙ !−1
401 Hz(cid:19)−1/3
(cid:18) νs
0.06 ME!" ( Mmin/ Mmax)α/3
M⊙(cid:19)−1/3 Mmax
1301/7
#
(9)
(10)
(cid:18) Rs
10 km(cid:19)(cid:18) M
M/M⊙ !−1
and
νs ≥ 1.0 × 10−13η I/1045 g cm2
10 km(cid:19)3/2(cid:18) M
(cid:18) Rs
M⊙(cid:19)−1/2 Mmax
0.06 ME! .
The maximum observed 2 -- 30 keV flux of SAX J1808.4 -- 3658 at the peak of the outburst was ≃ 3 ×
109 erg cm−2 s−1 (Cui et al. 1998), which corresponds to a 2 -- 30 keV luminosity of ≃ 0.03(Ωe/4π)LE,
where Ωe is the solid angle of emission and LE = 2×1038 erg s−1 is the Eddington critical luminosity
that corresponds to ME. However, the photon spectrum of SAX J1808.4 -- 3658 over most of the
RXTE 2 -- 100 keV bandpass is well described by a power-law of the form dN/dE ∝ E−2, where E
is the photon energy. The energy spectrum is thus very flat, νLν ∝ E0. As a result, the bolometric
luminosity of the source must be larger than estimated above and is probably about twice the
2 -- 30 keV luminosity, given the expected upper- and lower-energy cut-offs of the spectrum (see also
Heindl & Smith 1998 for spectral fits over the entire RXTE bandpass that show this effect). We
therefore adopt
Mmax = 0.06 ME.
Given a measurement of the spin frequency derivative in SAX J1808.4 -- 3658, we could use
relations (9) and (10) to place additional constraints on α for inner disk models or M and R for
neutron stars, similar to the discussion above for the presence of coherent pulsations (Figure 1).
However, variations of the spin frequency were not detected during the peak of the 1998 April/May
outburst of SAX J1808.4 -- 3658, leading to an upper limit on the spin frequency derivative of νs <
7 × 10−13 Hz s−1 (Chakrabarty & Morgan 1998). This measured upper limit is consistent with our
predicted bounds on the spin frequency derivative (9) and (10) if η ∼< 7.
3. THE MAGNETIC FIELD STRENGTH OF SAX J1808.4 -- 3658
In their discovery paper for SAX J1808.4 -- 3658, Wijnands & van der Klis (1998a) applied
"standard" magnetospheric disk accretion theory (essentially the 1G model discussed in §2) and
used the absence of centrifugal inhibition of accretion during the decline of the 1998 outburst to
infer that B ∼< (2−6)×108 G. As the outburst declined, pulsations continued to be detected, leading
to revised upper limits on the magnetic field strength [B ∼< (0.4 − 1.3) × 108 G, Cui et al. 1998;
B ∼< few ×107 G, Gilfanov et al. 1998]. However, as we showed in §2.2, simple scaling arguments
or models of the form (3) for the asymptotic regions of gas-pressure -- dominated accretion disks
cannot easily account for the coherent pulsations observed from SAX J1808.4 -- 3658 throughout its
1998 outburst. Therefore, the previous estimates of the magnetic field strength of SAX J1808.4 --
3658 which are based on similar scalings, are not valid. A more careful calculation is necessary
-- 9 --
to infer the magnetic field strength in this system. Because no spin frequency derivative has been
detected yet from this source (Chakrabarty & Morgan 1998), we cannot uniquely determine the
stellar magnetic field strength, but can only constrain it.
We begin with expression (2) for the interaction radius, which is derived directly from the
angular momentum equation (1), and impose the same constraints that led to equations (4) and
(5). We set the maximum and minimum mass accretion rates at which coherent pulsations were
detected to Mmax ≥ 0.06 ME and Mmin ≤ Mmax/130 as inferred from observations (see Cui et
al. 1998 and §2.3). We assume a neutron star radius in the 10 -- 15 km range, as predicted by
current equations of state (see Fig. 1). We also assume a neutron star mass in the 1.4 -- 2.3 M⊙
range, consistent with the inferred masses of recycled millisecond pulsars (Thorsett & Chakrabarty
1999), the upper limits on neutron star masses in LMXBs inferred from kHz QPO observations
(Miller et al. 1998; see also Zhang, Strohmayer, & Swank 1997), and the possible neutron star mass
measurement in 4U 1820−30 (Zhang et al. 1998a).
The limits on the magnetic field strength of SAX J1808.4 -- 3658 also depend upon the allowed
range for the parameter γB, which in turn depends upon the fractional width ∆r0/r0 of the bound-
ary layer and the magnetic pitch Bφ/Bp within this layer (see Ghosh & Lamb 1991 for a detailed
discussion). The fractional width should be significantly smaller than unity in order for the bound-
ary layer approximation (2) to be valid. Assuming a dipolar poloidal field and neglecting mass
loss from the disk as well as toroidal screening currents in both the disk and the magnetosphere,
one finds ∆r0/r0 ≃ 0.3 (Ghosh & Lamb 1991; Daumerie 1996). However, relaxing any of these
assumptions leads to a significantly narrower boundary layer. We therefore conservatively adopt
0.01 ∼< ∆r/r0 < 1.
The toroidal component Bφ of the magnetic field is produced by the differential rotation
of gas in the accretion disk with respect to the stellar spin. The magnetic energy stored in the
magnetosphere increases with the twisting of the field lines (see Zylstra 1988 and references therein).
There is an upper bound on this energy (and hence on the twisting of the field lines and the magnetic
pitch) for a semi-infinite space in which the magnetic field strength goes to zero at infinity (Aly
1984, 1991; Zylstra 1988).
It has been argued that, above this upper bound, the only existing
configurations are those with field lines that close at infinity (i.e., open field lines; Aly 1984, 1991).
Simple estimates of the maximum twisting of the field lines (see Ghosh & Lamb 1991 and reference
therein) as well as detailed numerical calculations of the magnetospheric structure (Zylstra 1988)
lead to maximum values of Bφ/Bp ∼ 1. Combining this with our constraints on ∆r0/r0, we
therefore adopt 0.01 ∼< γB( M ) ∼< 1.
Applying all these constraints to SAX J1808.4 -- 3658, we obtain
µ ∼> 0.3 × 1026hγB( Mmax)i−1/2(cid:18) M
1.4M⊙(cid:19)1/4(cid:18) Rs
10 km(cid:19)9/4 Mmax
0.06 ME!1/2
G cm3
(11)
-- 10 --
and
2. -- Constraints on the
Fig.
magnetic dipole moment µ of
SAX J1808.4 -- 3658 as a function of
boundary layer parameter γB( M ).
The upper limit on µ comes from
requiring that accretion is not cen-
trifugally inhibited at the minimum
M at which coherent pulsations
were detected. The lower limit on
µ comes from requiring that the ac-
cretion disk is terminated above the
stellar surface at the maximum M
at which coherent pulsations were
detected. Note that γB depends on
M , and so must have different val-
ues at the minimum and maximum
M considered here. Conservatively,
0.01
< γB < 1.
∼
µ ∼< 10 × 1026" γB( Mmin)
2.3M⊙(cid:19)5/6(cid:18) Rs
15 km(cid:19)1/2
(cid:18) M
4.6 × 10−4 ME!1/2
401 Hz(cid:19)−7/6
(cid:18) νs
Mmin
0.01
#−1/2
G cm3 ,
(12)
which correspond to a dipolar magnetic field of a few times 108 G at the stellar pole. Figure 2
shows the limits on the magnetic dipole moment as a function of the (unknown) parameter γB( M ).
Finally, we can estimate the mass accretion rate Meq for which the observed spin frequency of
the neutron star is the equilibrium frequency (at which the accretion torque is zero). The result is
Meq = 2 × 10−11ω−7/3
c
(cid:18) γB
0.1(cid:19)(cid:18)
µ
1026 G cm3(cid:19)2(cid:18) M
1.4M⊙(cid:19)−5/3(cid:18) νs
401 Hz(cid:19)7/3
M⊙ yr−1 ,
(13)
Meq depends
where ωc is the critical fastness parameter (Ghosh & Lamb 1979b). The value of
sensitively on the (unknown) magnetic field strength and on γB( M ). However, for values of these
Meq agrees with the long-term average
parameters consistent with the constraints imposed above,
accretion rate of SAX J1808.4 -- 3658 inferred by Chakrabarty & Morgan (1998) from the fluence of
the X-ray outburst and the outburst recurrence time. Moreover, in order for an LMXB with a 2 hr
binary period to be below the separatrix between transient and persistent systems in the diagram
M ∼< 3 × 10−11 M⊙ yr−1 (van Paradijs 1996;
of donor mass versus orbital period, it must have
King, Kolb, & Szuszkewicz 1997), also consistent with the value estimated above (Chakrabarty &
Morgan 1998).
-- 11 --
4. DISCUSSION
In this paper we have studied the disk-magnetosphere interaction in the first known accretion-
powered, millisecond pulsar, SAX J1808.4 -- 3658. We have demonstrated that various simple models
of inner accretion disks are not consistent with observations. We have also argued that the magnetic
field strength of SAX J1808.4 -- 3658 is a few times 108 G at the stellar pole, using very general
constraints on the properties of the neutron star and on the physics of the inner accretion disk flow.
In this section, we discuss the implications of our results for models of the disk-magnetosphere
interaction around weakly magnetic neutron stars. We also compare SAX J1808.4 -- 3658 to the
recycled millisecond radio pulsars and to the non-pulsing neutron-star LMXBs.
4.1. The Disk-Magnetosphere Interaction in Weakly Magnetic Neutron Stars
BATSE observations of strongly magnetic (B ∼ 1012 G) accretion-powered pulsars in Be/X-ray
transients have shown that the scaling of the radius r0 of interaction between the stellar magnetic
field and a gas-pressure dominated disk flow can account for the observations (Ghosh 1996; Finger,
Wilson, & Harmon 1996; Bildsten et al. 1997). However, in §2 we showed that this same scaling
is inconsistent with the detection of coherent pulsations throughout the 1998 April/May outburst
of SAX J1808.4 -- 3658, for most equations of state of neutron star matter. This is not surprising,
given the very different physical conditions in the interaction regions around weakly- and strongly-
magnetic neutron stars.
In writing the stress balance equation (1) and the scaling (2), we made a number of implicit
assumptions regarding the properties of the neutron star and the inner accretion disk flow: first,
that the stellar magnetic field is dipolar; second, that the pulsar magnetic moment is parallel to its
spin axis; third, that the dominant mechanism for removing angular momentum from the accreting
gas in the disk is magnetic stress; and finally, that the gravitational field is Newtonian everywhere
and hence that stable, circular Keplerian orbits exist at all radii. Although these assumptions are
valid (or at least are reasonable approximations) in the case of a strongly magnetic neutron star,
they break down when r0 is comparable to the stellar radius (see Lai 1998 for solutions of the
structure of the inner accretion disk where some of these effects have been taken into account in a
simple way).
Let us examine the validity of each of these assumptions. First, magnetic stresses would remove
angular momentum faster than predicted by equation (2) if higher-order multipoles were present
in the pulsar magnetic field. For example, if the stellar magnetic field could be described entirely
by a multipole of order l (i.e, the magnetic field strength in the equator was B = µl/rl+1), then
the Keplerian frequency at r0 would be
ν0 = (2π)−1γ−3/(4l−1)
B
(GM )(2l+1)/(4l−1)µ−6/(4l−1) M 3/(4l−1) .
(14)
Equation (14) shows that, if γB depends only weakly on M , the dependence of ν0 on M weakens
-- 12 --
Fig. 3. -- The effect of the presence of a
quadrupole magnetic moments on the con-
straints on the neutron star mass and ra-
dius in SAX J1808.4 -- 3658. The solid lines
represent mass-radius relations of stable
neutron stars for various proposed equa-
tions of state (Cook et al. 1994). The
dashed lines represent the limits imposed
by the 1G model for the inner accretion
disk and for different fractional contribu-
tions of an aligned quadrupole moment to
the magnetic field of the neutron star. The
allowed region for each model is to the left
of the corresponding dashed curve.
as the order l of the multipole increases: for a dipolar (l = 2) field ν0 ∝ M 0.43, for a quadrupolar
(l = 4) field ν0 ∝ M 0.2, etc. As a result, the existence of intrinsic or induced multipole moments in
the magnetosphere can significantly weaken the M -dependence of ν0, hence making the scaling (14)
consistent with observations of SAX J1808.4 -- 3658. As an illustration of this, Figure 3 shows the
constraints imposed by the 1G model on the mass and radius of the neutron star in SAX J1808.4 --
3658 for various contributions of an aligned quadrupole moment to the pulsar magnetic field.
Clearly, even a modest quadrupole moment (which would be completely negligible at distances
larger than a few stellar radii from the surface) is enough to make the 1G model consistent with
observations.
The intrinsic multipole moments in neutron star LMXBs are thought to be weaker than their
dipole moments, based on the current spin frequencies and spin frequency derivatives of the millisec-
ond radio pulsars thought to be their descendants (Arons 1993). However, electrical currents in the
accretion disk and the magnetosphere can significantly enhance the strength of multipole moments
by altering the magnetic field topology in the interaction region (see, e.g., Ghosh & Lamb 1979a).
Moreover, electrical currents on the neutron star induced by the presence of an accretion disk can
produce multipoles as strong as the intrinsic dipole, provided that the disk extends very close to
the stellar surface (Psaltis, Lamb, & Zylstra 1996). The corotation radius rco = (GM/4π2ν2
s )1/3
(which is an upper bound on r0 when coherent pulsations are detected; see, e.g., Ghosh & Lamb
1979a) is only ≃ 2.8 times the neutron star radius in SAX J1808.4 -- 3658, and therefore none of the
above effects are negligible.
Second, if any of the magnetic moments of SAX J1808.4 -- 3658 is not aligned to the stellar
rotation axis (which is actually a necessary condition for the system to appear as a pulsar) then
the scaling of the inner Keplerian disk radius with accretion rate may be different than what is
-- 13 --
Fig.
4. -- Inferred pulsar dipole mag-
netic field strength at the stellar equator,
as a function of spin period. The small
dots are normal radio pulsars (Taylor et
al. 1993). The open circles are millisec-
ond radio pulsars which are not members
of globular clusters, and for which the in-
ferred field strengths have been corrected
for the apparent period derivative caused
by their transverse velocities (Camilo et al.
1994). The error bar shows the allowed
range of field strengths for SAX J1808.4 --
3658. The spin-up line (solid ) is shown
here for illustrative reasons only and rep-
resents the minimum period of recycled
pulsars for a given field strength, assum-
ing a specific model for the inner accretion
disk and the disk-magnetosphere interac-
tion. The death line (dashed ) is an em-
pirical estimate of the maximum period at
which pulsars have detectable radio emis-
sion for a given field strength.
predicted by the models considered above. This case has never been treated in the literature
because it corresponds to a time-dependent non-steady-state problem. It is beyond the scope of
the current paper to generalize models of the interaction between the pulsar magnetic field and the
accretion disk to time-dependent situations. We will therefore not consider this case any further.
Third, when the accretion disk penetrates very close to the neutron star, the effect of magnetic
stresses is amplified by radiation drag forces that can efficiently remove angular momentum from
the accreting gas (Miller & Lamb 1996; Miller et al. 1998). Finally, the specific angular momentum
of gas at radii comparable to the innermost stable orbit around a compact object has a weaker
dependence on radius than estimated in a Newtonian gravitational field.
All the effects discussed above result in a weaker
M -dependence of ν0, thus making scaling (2)
consistent with observations of SAX J1808.4 -- 3658.
4.2. Comparison with Millisecond Radio Pulsars
When mass transfer onto the neutron star ends, SAX J1808.4 -- 3658 may appear as a millisecond
radio pulsar (Wijnands & van der Klis 1998a; see also Bhattacharya & van den Heuvel 1991).
Figure 3 compares the current spin frequency and inferred magnetic field strength of SAX J1808.4 --
3658 at the stellar equator with those of the known normal and millisecond radio pulsars (Taylor,
Manchester, & Lyne 1993). We only include millisecond radio pulsars in the Galactic disk, since the
-- 14 --
inferred spin frequency derivatives (and hence the magnetic field strengths) of pulsars in globular
clusters are probably contaminated by gravitational acceleration in the potential of the cluster.
Moreover, we use magnetic field strengths that have been corrected for the effect of the apparent
spin frequency derivative caused by the non-negligible transverse velocities of the pulsars (Camilo,
Thorsett, & Kulkarni 1994; Toscano et al. 1999). Note, however, that the inferred magnetic dipole
moments assume orthogonal dipole rotators, and model-dependent systematic effects have not been
taken into account.
The mass of the companion of SAX J1808.4 -- 3658 is ∼< 0.3M⊙, and hence the binary must be
close to the termination of its X-ray phase (Chakrabarty & Morgan 1998). The current magnetic
field strength and spin frequency of the neutron star should thus be very similar to that of the
descendant millisecond radio pulsar. Figure 3 shows that the descendant of SAX J1808.4 -- 3658
should appear as a normal millisecond radio pulsar in the Galactic disk. Given the small number
statistics and the uncertainty in our estimate of the magnetic field strength in SAX J1808.4 -- 3658
and in the millisecond radio pulsars, it is not possible to determine whether the field strength
of SAX J1808.4 -- 3658 is lower or higher than those in the known millisecond radio pulsars. Such
a determination would be of interest in comparing SAX J1808.4 -- 3658 to the other non-pulsing
LMXBs.
4.3. Comparison with Non-Pulsing LMXBs
SAX J1808.4 -- 3658 is the only known weakly-magnetic accreting neutron star with persistent
coherent pulsations at the neutron star spin frequency. Upper limits on the amplitudes of periodic
oscillations in other neutron-star LMXBs range from ∼< 1% for the most luminous Z sources to a
few percent for the less luminous atoll sources (see Vaughan et al. 1994 and references therein).
Coherent millisecond oscillations, however, possibly at the spin frequencies of the neutron stars,
have been observed in several neutron-star LMXBs during Type I X-ray bursts (Strohmayer, Swank,
& Zhang 1998).
The absence of persistent coherent pulsations in most LMXBs might be due to neutron star
magnetic field strengths that are significantly different than that in SAX J1808.4 -- 3658. Specifically,
the field strengths in the non-pulsing LMXBs may be so strong that accretion is centrifugally
inhibited, or so weak that the accretion disk extends all the way to the neutron star surface.
Alternatively, the neutron stars in the non-pulsing LMXBs may be surrounded by optically thick
scattering media that spread the pulsar beams and attenuate the oscillations below present detection
thresholds (Brainerd & Lamb 1987; Kylafis & Phinney 1989). We consider here each possibility
separately.
It is rather unlikely that the field strengths of all non-pulsing LMXBs are so strong that accre-
tion is centrifugally inhibited, since at least some are accreting at near-Eddington rates. Moreover,
if the field strengths were all so strong, then the observational fact that all known millisecond radio
-- 15 --
pulsars lie below the so-called spin-up line in the B-P diagram (Fig. 3) would have to be a selection
effect or a coincidence. This is because the spin-up line defines the minimum spin period of a
recycled pulsar (which is the equilibrium period at the Eddington accretion rate) for a given field
strength, provided that accretion is not centrifugally inhibited.
On the other hand, if the field strengths in the non-pulsing LMXBs are significantly weaker
than that of SAX J1808.4 -- 3658, then a significant fraction of the gas in the accretion disk may
be reaching the stellar surface in the disk plane, producing undetectable periodic oscillations at
the neutron star spin frequency. This is apparently consistent with identifying the frequencies of
the observed kHz QPOs in the non-pulsing LMXBs with Keplerian orbital frequencies in the inner
accretion disk (van der Klis et al. 1996; Strohmayer et al. 1996; Miller et al. 1998).
However, there are at least two other transient LMXBs (Aql X-1 and 4U 1608−52) whose
outburst histories, X-ray spectral properties, and broad-band noise features are similar to those of
SAX J1808.4 -- 3658 (see, e.g., Heindl & Smith 1998; Wijnands & van der Klis 1998b), and yet do
not show periodic pulsations in their persistent emission. At the same time, these systems do show
kHz QPOs and have other timing and spectral properties characteristic of the non-pulsing LMXBs
(Zhang, Yu, & Zhang 1998b; Yu et al. 1997). In particular, the X-ray spectrum of 4U 1608−52,
which is a direct probe of the structure and dynamics of the inner accretion disk flow, is nearly
identical to that of SAX J1808.4 -- 3658 (Heindl & Smith 1998), with the same power-law index and
high-energy cut-off. It seems unlikely that the X-ray spectra of these two sources could be so similar
if their field strengths (and thus their inner disk radii) differ significantly. This is particularly true
of the spectra at high energies (20 -- 100 keV), which must be formed close to the neutron star surface
and are thus highly sensitive to the inner disk flow geometry. In fact, if the high-energy spectra are
produced by Comptonization of high-harmonic self-absorbed cyclotron photons (Psaltis et al. 1995),
then the stellar field strengths can be inferred from the hardness of the 2 − 20 keV X-ray spectra
(Psaltis 1998; Psaltis & Lamb 1998). Indeed, the magnetic field strengths of the low-luminosity
atoll sources, such as Aql X-1 and 4U 1608−52, have been inferred to be ∼ 5 × 108 G (see Psaltis
& Lamb 1998), i.e., very similar to the magnetic field strength of SAX J1808−36 we estimated in
§3.
If the field strength in SAX J1808.4 -- 3658 is comparable to that of some non-pulsing LMXBs,
then the absence of periodic oscillations in the persistent emission from most LMXBs must be
attributed to another, external effect. The presence of a scattering medium around the neutron
star and its magnetosphere has been invoked previously as an explanation of the non-detection of
periodic oscillations in neutron-star LMXBs (Brainerd & Lamb 1987; Kylafis & Phinney 1989). The
high (τ ∼ 4 − 5) optical depths inferred from X-ray spectra of the Z and atoll sources are consistent
with this hypothesis (see Psaltis et al. 1995 and references therein). However, RXTE/HEXTE
observations of SAX J1808.4 -- 3658 indicate an optical depth of τ = 4.0 ± 0.2, based on the slope
and high-energy cutoff of the X-ray spectrum, at the same time that pulsations were detected
(Heindl & Smith 1998; Heindl 1998, private communication; note, however, that Gilfanov et al.
[1998] analyzed the same data and reached a different conclusion.). This is very similar to the
-- 16 --
optical depth inferred for the non-pulsing source 4U 1608−52 (Heindl & Smith 1998), which we
argued above has a comparable magnetic field strength. Thus, if the optical depth of SAX J1808.4 --
3658 is indeed ∼> 3, then the reason coherent pulsations are detected only from this LMXB cannot
simply be weaker attenuation of the beaming oscillation in the scattering medium.
A number of additional effects may be responsible for the detectability of coherent pulsations
in SAX J1808.4 -- 3658. For example, its very small measured mass function is suggestive of a nearly
face-on viewing angle of the binary (Chakrabarty & Morgan 1998). In this case, the pulsed emission
may propagate through a region of lower optical depth close to the pulsar rotation poles, whereas
the majority of the (non-pulsed) X-ray photons may travel through optically thick regions nearby.
However, it is important to note that the presence of a shallow partial eclipse in SAX J1808.4 -- 3658
has been suggested (Chakrabarty & Morgan 1998; Heindl & Smith 1998). Confirmation of this
feature would rule out a face-on orientation of the binary, thus challenging the above explanation.
Alternatively, different stellar field topologies in SAX J1808.4 -- 3658 and the non-pulsing neutron
star LMXBs, possibly related to different prior evolution or even to different magnetic inclinations,
may be responsible.
A number of other LMXBs share many characteristics of SAX J1808.4 -- 3658 and are thus
good candidates for detecting periodic pulsations in their persistent emission. For example, as
we discussed above, Aql X-1 and 4U 1608−52 are transients with X-ray spectra very similar to
SAX J1808.4 -- 3658, suggesting similar field strengths and inner disk flows. The nearly sinusoidal
optical orbital-phase lightcurves of three other LMXBs (4U 1636−53, GX 9+9, and 4U 1735−44)
suggest that these are also viewed nearly face-on (see, van Paradijs & McClintock 1995). Finally,
GS 1826−26 (Homer, Charles, & O'Donoghue 1998) and MS 1603+2600 (Hakala et al. 1998) are
other LMXBs with low-mass companions in ∼ 2-hour orbits, in which mass transfer is probably
driven by angular momentum losses due to gravitational radiation. These binaries are thus in
an evolutionary stage very similar to that of SAX J1808.4 -- 3658.
In particular, GS 1826−26 is
a transient X-ray burster with a low-amplitude modulation of its optical brightness, again possi-
bly suggesting a face-on orientation (Homer et al. 1998). Detecting or imposing stringent upper
limits on coherent pulsations in the persistent emission of any of these sources will be crucial in
understanding the characteristics of SAX J1808.4 -- 3658 and the relationship between LMXBs and
millisecond radio pulsars.
We are grateful to W. Heindl and Wei Cui for sharing their results prior to publication. We
also thank Steve Thorsett for pointing out some recent work on the magnetic fields of recycled
pulsars, Uli Kolb for discussions on the disk-instability model of X-ray transients and on the binary
properties of SAX J1808.4 -- 3658, as well as Lars Bildsten, Vicky Kalogera, Michiel van der Klis,
and Rudy Wijnands for many useful discussions and for carefully reading the manuscript. This
work was supported in part by a postdoctoral fellowship of the Smithsonian Institution (D. P.), by
a NASA Compton GRO Postdoctoral Fellowship at MIT under grant NAG 5-3109 (D. C.), and
by several NASA grants under the RXTE Guest Observer Program. We are also grateful for the
-- 17 --
hospitality of the Astronomy Group at the University of Leicester and the Astronomical Institute
at the University of Amsterdam, where parts of this work were completed.
REFERENCES
Aly, J.-J. 1984, ApJ, 283, 349
-- -- -- . 1991, ApJ, 375, L61
Arons, J. 1993, ApJ, 408, 160
Alpar, M. A., & Shaham, J. 1985, Nature, 316, 239
Bhattacharya, D., & van den Heuvel, E. P. J. 1991, Phys. Rep., 203, 1
Bildsten, L. et al. 1997, ApJS, 113, 367
Brainerd, J., & Lamb, F. K. 1987, ApJ, 317, L33
Burderi, L., & King, A. R. 1998, ApJ, 505, L135
Camilo, F., Thorsett, S. E., & Kulkarni, S. R. 1994, ApJ, 421, L15
Chakrabarty, D. et al. 1997a, ApJ, 474, 414
-- -- -- . 1997b, ApJ, 481, L101
Chakrabarty, D., & Morgan, E. H. 1998b, Nature, 394, 346
Cook, G., Shapiro, S. L., & Teukolsky, S. A. 1994, ApJ, 424, 823
Cui, W., Morgan, E. H., & Titarchuk, L. 1998, ApJ, 504, L27
Davidson, K., & Ostriker, J. P. 1973, ApJ, 179, 585
Daumerie, P. 1996, Ph. D. thesis, University of Illinois at Urbana-Champaign
Finger, M. H., Wilson, R. B., & Harmon, B. A. 1996, ApJ, 459, 288
Ghosh, P. 1996, ApJ, 459, 244
Ghosh, P., & Lamb, F. K. 1978, ApJ, 223, L83
-- -- -- . 1979a, ApJ, 232, 259
-- -- -- . 1979b, ApJ, 234, 296
-- -- -- . 1991 in Neutron Stars: Theory and Observation, ed. J. Ventura and D. Pines, (Dordrecht:
Kluwer), p. 363
-- -- -- . 1992 in X-ray Binaries and Recycled Pulsars, ed. E. P. J. van den Heuvel and S. A. Rappa-
port (Dordrecht: Kluwer), 487
Giacconi, R., Gursky, H., Kellogg, E., Levinson, R., Schreier, E., & Tananbaum,, H. 1971, ApJ,
167, L67
Gilfanov, M., Revnivtsev, M., Sunyaev, R., & Churazov, E. 1998, A&A, 338, L83
-- 18 --
Hakala, P. J., Chaytor, D. H., Vilhu, O., Pirrola, V., Morris, S. L., & Muhli, P. 1998, A&A, 333,
540
Heindl, W. A., & Smith, D. M. 1998, ApJ, 506, L35
Homer, L., Charles, P. A., & O'Donoghue, D. 1998, MNRAS, 298, 497
in't Zand, J. J., Heise, J., Muller, J. M., Bazzano, A., Cocchi, M., Natalucci, L., & Ubertini, P.
1998, A&A, 331, L25
King, A. R., Kolb, U., & Szuszkiewicz, E. 1997, ApJ, 488, 89
Kylafis, N., & Phinney, E. S. 1989, in Timing Neutron Stars, eds. H. Ogelman and E. P. J. van den
Heuvel (Dordrecht: Kluwer), 731
Lai, D. 1998, ApJ, 502, 721
Lamb, F. K., Pethick, C. J., & Pines, D. 1973, ApJ, 184, 271
Lamb, F. K., Shibazaki, N., Alpar, M. A., & Shaham, J. 1985, Nature, 317, 681
Lewin, W. H. G., van Paradijs, J., & Taam, R. 1995, in X-ray Binaries, ed. W. H. G. Lewin, E. P. J.
van den Heuvel, & J. van Paradijs, (Cambridge: University Press), 175
Miller, M. C., & Lamb, F. K. 1996, ApJ, 470, 1033
Miller, M. C., Lamb, F. K., & Psaltis, D. 1998, ApJ, 508, 791
Nelson, R. et al. 1998, ApJ, 488, L117
Pringle, J. E., & Rees, M. J. 1972, A&A, 21, 1
Psaltis, D. 1998, Ph. D. thesis, University of Illinois at Urbana-Champaign
Psaltis, D., Belloni, T., & van der Klis, M. 1999a, ApJ, submitted
Psaltis, D., & Lamb, F. K. 1998, in Neutron Stars and Pulsars, ed. N. Shibazaki, N. Kawai, S.
Shibata, & T. Kifune (Tokyo: Universal Academy Press), p. 179
Psaltis, D., Lamb, F. K., & Miller, G. S. 1995, ApJ, 454, L137
Psaltis, D., Lamb, F. K., & Zylstra, G. J. 1996, Proceedings of the NATO ASI: Solar and Astro-
physical MHD Flows, Astr. Lett. & Comm., 34, 377
Psaltis, D., Wijnands, R., Homan, J., Jonker, P., van der Klis, M., Miller, M. C., Lamb, F. K.,
Kuulkers, E., van Paradijs, J., & Lewin, W. H. G. 1999b, ApJ, submitted
Strohmayer, T. E., Swank, J.., & Zhang, W. 1998, Nucl. Phys. B (Proc. Suppl.), 69, 129
Strohmayer, T. E., Zhang, W., Swank, J. H., Smale, A., Titarchuk, L., Day, C., & Lee, U. 1996,
ApJ, 469, L9
Taylor, J. H., Manchester, R. N., & Lyne, A. G. 1993, ApJS, 88, 529
Thorsett, S. E., & Chakrabarty, D. 1999, ApJ, 512, in press (astro-ph/9803260)
Torkelsson, U. 1998, MNRAS, 298, L55
-- 19 --
Toscano, M., Sandhu, J. S., Bailes, M., Manchester, R. N., Britton, M. C., Kulkarni, S. R., Anderson,
S. B., & Stappers, B. W. 1999, MNRAS, submitted (astro-ph/9811398)
Treves, A., Maraschi, L., & Abramowicz, M. 1988, PASP, 100, 427
van der Klis, M. 1989, ARA&A, 27, 517
-- -- -- . 1998, in The Many Faces of Neutron Stars, ed. R. Buccheri, J. van Paradijs, & M. A. Alpar
(Dordrecht: Kluwer), 337
van der Klis, M., Jansen, F., van Paradijs, J., van den Heuvel, E. P. J., & Lewin, W. H. G. 1985,
Nature, 316, 225
van der Klis, M., Swank, J. H., Zhang, W., Jahoda, K., Morgan, E. H., Lewin, W. H. G., Vaughan,
B., & van Paradijs, J. 1996, ApJ, 469, L1
van Kerkwijk, M. H., Chakrabarty, D., Pringle, J. E., & Wijers, R. A. M. J. 1998, ApJ, 499, L27
van Paradijs, J. 1996, A&A, 464, L139
van Paradijs, J., & McClintock, J. 1995, in X-ray Binaries, ed. W. H. G. Lewin, E. P. J. van den
Heuvel, & J. van Paradijs, (Cambridge: University Press), 58
Vaughan, B. A., van der Klis, M., Wood, K. S., Norris, J. P., Hertz, P., Michelson, P. F., van
Paradijs, J., Lewin, W. H. G., Mitsuda, K., & Penninx, W. 1994, ApJ, 435, 362
Wang, Y.-M. 1996, ApJ, 465, L111
White, N. E., Nagase, F., & Parmar, A. N. 1995, in X-ray Binaries, ed. W. H. G. Lewin, J. van
Paradijs, and E. P. J. van den Heuvel (Cambridge: Cambridge University Press), 1
Wijnands, R., & van der Klis, M. 1998a, Nature, 394, 344
-- -- -- . 1998b, ApJ, 507, L63
Yi, I., & Wheeler, J. G. 1998, ApJ, 498, 802
Yu, W. et al. 1997, ApJ, 490, L153
Zhang, W., Smale, A. P., Strohmayer, T. E., & Swank, J. H. 1998a, ApJ, 500, L171
Zhang, W., Strohmayer, T. E., & Swank, J. H. 1997, ApJ, 482, L167
Zhang, S. N., Yu, W., & Zhang, W. 1998b, ApJ, 494, L71
Zylstra, G. J. 1988, Ph. D. thesis, University of Illinois at Urbana-Champaign
This preprint was prepared with the AAS LATEX macros v4.0.
-- 20 --
Table 1. Scaling of the Keplerian Frequency at the Coupling Radiusa
Disk Modelb
νK
,
0 (Hz)
1G
1R
2B
2S
430
210
80
50
α
0.38
0.23
0.72
2.55
β
−0.87
−0.77
−0.86
−1.20
γ
0.82
0.70
0.43
−0.60
aAfter Ghosh & Lamb 1992.
b1G: Optically thick, gas-pressure dominated (GPD) disk; 1R: Optically thick,
radiation-pressure dominated (RPD) disk; 2B: Two-temperature, optically thin
GPD disk with Comptonized bremsstrahlung; 2S: Two-temperature, optically
thin GPD disk with Comptonized soft photons (for references see Ghosh & Lamb
1992).
|
astro-ph/0611534 | 1 | 0611 | 2006-11-16T18:44:27 | A non-axisymmetric magnetorotational instability of a purely toroidal magnetic field | [
"astro-ph"
] | We consider the flow of an electrically conducting fluid between differentially rotating cylinders, in the presence of an externally imposed toroidal field B_0 (r_i/r) e_phi. It is known that the classical, axisymmetric magnetorotational instability does not exist for such a purely toroidal imposed field. We show here that a non-axisymmetric magnetorotational instability does exist, having properties very similar to the axisymmetric magnetorotational instability in the presence of an axial field. | astro-ph | astro-ph | A Non-axisymmetric Magnetorotational Instability of a Purely
Toroidal Magnetic Field
Rainer Hollerbach1 and Gunther Rudiger, Manfred Schultz, D. Elstner2
1Department of Applied Mathematics,
University of Leeds, Leeds, LS2 9JT, United Kingdom
2Astrophysikalisches Institut Potsdam,
An der Sternwarte 16, D-14482 Potsdam, Germany
(Dated: August 16, 2018)
Abstract
We consider the flow of an electrically conducting fluid between differentially rotating cylinders,
in the presence of an externally imposed toroidal field B0(ri/r) eφ. It is known that the classical,
axisymmetric magnetorotational instability does not exist for such a purely toroidal imposed field.
We show here that a non-axisymmetric magnetorotational instability does exist, having properties
very similar to the axisymmetric magnetorotational instability in the presence of an axial field.
PACS numbers: 47.20.-k, 47.65.+a, 95.30.Qd
6
0
0
2
v
o
N
6
1
1
v
4
3
5
1
1
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
The magnetorotational instability (MRI) is a mechanism whereby a hydrodynamically
stable differential rotation flow may be destabilized by the addition of a magnetic field. An
obvious question then is whether different orientations of the field yield different types of
instability, or the same, or none at all. The original view was that the axial component of
the field is the only important one, with any azimuthal component playing no essential role,
and incapable of producing any instabilities on its own [1, 2]. This view was altered by the
discovery that a mixed axial and azimuthal field yields instabilities quite different in many
ways from those found with a purely axial field [3, 4]. In this letter we show that even a
purely azimuthal field yields an MRI, and compare its properties with the previously known
types.
While its most important application is to astrophysical accretion disks [2], the MRI was
originally discovered in the much simpler Taylor-Couette problem, consisting of the flow
between differentially rotating cylinders [1]. Because of its relative simplicity, this geometry
has proven particularly amenable both to theoretical analyses of the MRI [5, 6, 7], as well as
to its recent experimental realization [8, 9]. We too shall examine the MRI in this context.
Consider therefore an electrically conducting fluid confined between two concentric cylin-
ders of radii ri and ro, rotating at rates Ωi and Ωo, chosen to satisfy Ωor2
o > Ωir2
i . That is,
the angular momentum increases outward, so by the familiar Rayleigh criterion the flow is
hydrodynamically stable, with the angular velocity given by
Ω(r) = A + B/r2,
(1)
(2)
.
where
A =
o − Ωir2
Ωor2
o − r2
r2
i
i
, B =
i r2
r2
o(Ωi − Ωo)
o − r2
r2
i
In this work we will fix ro = 2ri and Ωo = Ωi/2, so A and B simplify to 1
3Ωi and 2
3 Ωir2
i ,
respectively. The essence of the MRI then is to ask whether the addition of a magnetic field
can destabilize this flow.
If the imposed field is purely axial, B0 = B0 ez, the profile (1) can be destabilized,
provided the rotation rates are sufficiently great, and the field strength B0 is neither too
weak nor too strong [5, 6]. Specifically, the magnetic Reynolds number Rm = Ωir2
i /η, where
η is the magnetic diffusivity, must exceed O(10), and the Lundquist number S = B0ri/η√ρµ,
where ρ is the density and µ the permeability, must be around 3 − 10.
2
In contrast, if the imposed field is mixed axial and azimuthal, B0 = B0 ez + βB0(ri/r) eφ,
where β is around 1− 10, then the profile (1) can again be destabilized, but at very different
rotation rates and field strengths [3, 4]. The relevant parameter measuring the rotation rates
is now not the magnetic Reynolds number, but rather the hydrodynamic Reynolds number
Re = Ωir2
i /ν, where ν is the viscosity. Similarly, the parameter measuring the field strength
is not the Lundquist number, but instead the Hartmann number Ha = Bori/√ρµην. The
MRI sets in when Re >
∼ O(103), and Ha ≈ O(10).
To compare these results, we note that the two sets of parameters are related by
Re = Rm Pm−1 and Ha = S Pm−1/2, where Pm = ν/η is the magnetic Prandtl number,
a material property of the fluid. Typical values for liquid metals are O(10−6). Translating
the results for the purely axial field, we thus obtain Re >
∼ O(107) and Ha ≈ O(104), both
several orders of magnitude greater than for the mixed field. It is perhaps not surprising
then that the MRI has been obtained experimentally for the mixed field [8, 9], but not (yet)
for the purely axial field [5].
As different as they are, one feature these two types of MRI have in common is that
they are both axisymmetric. Non-axisymmetric modes have also been explored, for both
the purely axial [7] as well as the mixed fields [4]. For a purely axial field the relevant
parameters are still Rm and S, but the critical values Rmc are somewhat larger than for
the axisymmetric modes, indicating that the axisymmetric MRI is the most unstable mode.
For mixed fields, one finds -- perhaps somewhat surprisingly -- that adding an azimuthal
component now has minimal effect, certainly far less than the reduction by four orders of
magnitude found for the axisymmetric modes. Evidently the relevant parameters continue
to be Rm and S, rather than Re and Ha.
What we wish to show in this letter then is that for these non-axisymmetric modes, one
can impose a purely azimuthal field, B0(ri/r) eφ, and still obtain an MRI, having all the
characteristics of the previous non-axisymmetric types of MRI [4, 7]. To this end, we solve
the linear stability equations
Rm
∂b
∂t
= ∇2b + ∇ × (u × B0) + Rm∇ × (U0 × b),
Re
∂u
∂t
= −∇p + ∇2u + Ha2 (∇ × b) × B0 + Re (U0 × ∇ × u + u × ∇ × U0),
(3)
(4)
where U0 = rΩ(r) eφ is the profile (1) whose stability we are exploring, and B0 = B0(ri/r) eφ
is the imposed azimuthal field. Length has been scaled by ri, time by Ω−1
, U0 and u as
i
3
riΩi, B0 as B0, and b as RmB0.
Taking the t, z and φ dependence to be of the form exp(σt + ikz + imφ), and using also
∇ · b = 0 to eliminate bz, the r and φ components of (3) become
Rm σbr = ∇2br − r−2br − 2imr−2bφ + imr−2ur − Rm imΩ br,
Rm σbφ = ∇2bφ − r−2bφ + 2imr−2br + imr−2uφ + 2r−2ur − Rm imΩ bφ + Rm Ω′rbr,
(5)
(6)
where primes denote d/dr. The components of (4) have a similar structure, but we will
not need to refer to them in the subsequent discussion, and hence do not list them. The
boundary conditions associated with (3) and (4) are
br = b′
φ + r−1bφ = ur = uφ = uz = 0
(7)
at r = ri and ro, corresponding to perfectly conducting, no-slip walls. The resulting one-
dimensional linear eigenvalue problem is solved by finite differencing in r, as in [7].
Figure 1 shows stability curves for m = 1, the only mode that appears to become unstable.
We see how an MRI exists that is remarkably similar to some of the results described above.
In particular, as Pm → 0, the relevant parameters are clearly once again Rm and S, with
the MRI arising if Rm >
∼ 80, and S ≈ 40 yielding the lowest value of Rmc. The specific
numbers are roughly an order of magnitude greater than for the axisymmetric MRI in the
axial field, but the basic scalings, and even the detailed shape of the instability curves, are
identical.
Figure 2 shows the real and imaginary parts of σ in the unstable regime. Remembering
that time has been scaled by Ω−1
i
, we see that we obtain growth rates as large as 0.05Ωi. So
again, while the particular number 0.05 is about an order of magnitude smaller than for the
axisymmetric MRI in the axial field, this non-axisymmetric MRI is clearly also growing on
the basic rotational timescale.
To understand why this non-axisymmetric MRI exists even for a purely toroidal field B0,
for which it is known that the axisymmetric MRI fails [1, 2], we need to consider the details
of (5) and (6). In particular, note that for m = 0, br completely decouples from everything
else, and inevitably decays away. Without br though, the MRI cannot proceed, as it relies
on the term Rm Ω′rbr in (6). In contrast, for m = 1, br is coupled both to bφ, coming from
∇2b, and to ur, from ∇ × (u × B0). And once br is coupled to the rest of the problem,
4
FIG. 1: The critical magnetic Reynolds number for the onset of the MRI, as a function of the
Lundquist number, for the different values of Pm indicated. m = 1, ro/ri = 2, Ωo/Ωi = 0.5.
FIG. 2: The real (top) and imaginary (bottom) parts of σ, as functions of S, with Rm fixed at 200.
The real part corresponds to the growth rate, the imaginary part to the azimuthal drift rate.
the term Rm Ω′rbr then allows the MRI to develop. Figure 3 presents an example of these
solutions, indicating how all three components of both u and b are indeed present.
There is however one aspect that is not entirely clear from this analysis, namely why
this non-axisymmetric MRI actually requires the term Rm Ω′rbr at all. In particular, the
axisymmetric MRI with a mixed field does not rely on it, but instead on the term 2r−2ur,
coming from ∇ × (u × B0) rather than Rm∇ × (U0 × b) [3]. The non-axisymmetric MRI
is evidently more like the axisymmetric MRI with a purely axial field, which also requires
the term Rm Ω′rbr, because in that case the term 2r−2ur is absent.
5
FIG. 3: The marginally stable solution at Pm = 0.1, S = 47.4 and Rm = 93 (see Fig. 1). From left
to right the r, φ and z components of u (top) and b (bottom). The real parts are solid, imaginary
parts dashed. Note how both u and b have the z components largest. The azimuthal wavenumber
is k = 1.88.
To summarize then, we have shown that even if the externally imposed magnetic field is
purely toroidal, one still obtains an MRI, simply non-axisymmetric rather than axisymmet-
ric. In other respects though this new instability is remarkably like the classical axisymmetric
MRI with a purely axial field, indeed far more like it than the axisymmetric MRI with a
mixed field, which yielded fundamentally different scalings. In contrast, the parameters here
continue to be Rm and S, just as in the classical MRI.
Note though that attempting to obtain this non-axisymmetric MRI in a laboratory ex-
periment would be even more difficult than attempting to obtain the classical axisymmetric
MRI in an axial field. First, the required rotation rates would be even greater, Re >
∼ O(108),
with all the difficulties that entails [10]. Even more daunting, imposing an azimuthal field of
the required strength, Ha ≈ O(105), would require a current along the central axis in excess
of 106 A, surely far beyond any feasible experiment.
6
This new non-axisymmetric MRI could have astrophysical applications though, since
many astrophysical objects do have predominantly azimuthal fields, in which case the re-
sults presented here suggest that this non-axisymmetric MRI could be preferred over the
axisymmetric MRI.
Despite the similarities in their fundamental scalings, the fact that the classical MRI is
axisymmetric, whereas this new MRI is non-axisymmetric, means the nonlinear equilibra-
tion, and hence the associated angular momentum transport, are potentially quite different,
which could again have astrophysical implications. Work on this is currently in progress.
Finally, one might ask how the results presented here change if one allows for a more
general toroidal field profile, B0 = (c1r−1 + c2r) eφ, where the term c2r corresponds to an
electric current flowing through the fluid itself, and not just along the central axis. Eq.
(4) then contains an additional term Ha2(∇ × B0) × b, which opens up the possibility of
instabilities driven entirely by this current ∇× B0, without any rotation necessarily present
at all [11]. Understanding how these current driven instabilities (also m = 1) interact with
the magnetically catalyzed but ultimately rotationally driven MRI presented here is also in
progress.
This work was supported by the German Leibniz Gemeinschaft, under program SAW.
[1] E. P. Velikhov, Sov. Phys. JETP 36, 995 (1959).
[2] S. A. Balbus and J. F. Hawley, Astrophys. J. 376, 214 (1991).
[3] R. Hollerbach and G. Rudiger, Phys. Rev. Lett. 95, art no 124501 (2005).
[4] G. Rudiger, R. Hollerbach, M. Schultz and D. Shalybkov, Astron. Nachr. 326, 409 (2005).
[5] H. T. Ji, J. Goodman, and A. Kageyama, Mon. Not. Roy. Astron. Soc. 325, L1 (2001).
[6] G. Rudiger and Y. Zhang, Astron. Astrophys. 378, 302 (2001).
[7] G. Rudiger, M. Schultz, and D. Shalybkov, Phys. Rev. E 67, art no 046312 (2003).
[8] F. Stefani, T. Gundrum, G. Gerbeth, G. Rudiger, M. Schultz, J. Szklarski and R. Hollerbach,
Phys. Rev. Lett. 97, art no 184502 (2006).
[9] G. Rudiger, R. Hollerbach, F. Stefani, T. Gundrum, G. Gerbeth and R. Rosner, Astrophys.
J. 649, L145 (2006).
[10] R. Hollerbach and A. Fournier, in AIP Conf. Proc. 733, MHD Couette Flows: Experiments
7
and Models, ed. R. Rosner, G. Rudiger and A. Bonanno, (New York: AIP), 114 (2004).
[11] R. J. Tayler, Mon. Not. Roy. Astron. Soc. 161, 365 (1973).
8
|
0710.4512 | 1 | 0710 | 2007-10-24T17:46:35 | Giant Radio Galaxies as a probe of the cosmological evolution of the IGM, I. Preliminary deep detections and low-resolution spectroscopy with the SALT | [
"astro-ph"
] | A problem of the cosmological evolution of the IGM is recalled and a necessity to find distant (z>0.5) giant radio galaxies (GRGs) with the lobe energy densities lower than about 10^{-14} J m^{-3} to solve this problem is emphasized. Therefore we undertake a search for such GRGs on the southern sky hemisphere using the SALT. In this paper we present a selected sample of the GRG candidates and the first deep detections of distant host galaxies, as well as the low-resolution spectra of the galaxies identified on the DSS frames. The data collected during the Performance Verification (P-V) phase show that 21 of 35 galaxies with the spectroscopic redshift have the projected linear size greater than 1 Mpc (for H_{0}=71 km\s\Mpc). However their redshifts do not exceed the value of 0.4 and the energy density in only two of them is less than 10^{-14} J m^{-3}. A photometric redshift estimate of one of them (J1420-0545) suggests a linear extent larger than 4.8 Mpc, i.e. a larger than that of 3C236, the largest GRG known up to now. | astro-ph | astro-ph |
Giant Radio Galaxies as a probe of the cosmological evolution of
the IGM, I. Preliminary deep detections and low-resolution
spectroscopy with the SALT
J. Machalski1, D. Kozie l-Wierzbowska1, M. Jamrozy1
1 Obserwatorium Astronomiczne, Uniwersytet Jagiello´nski,
ul. Orla 171, 30-244, Krak´ow, Poland
([email protected], [email protected], [email protected])
Abstract
A problem of the cosmological evolution of the IGM is recalled and a neces-
sity to find distant (z > 0.5) "giant" radio galaxies (GRGs) with the lobe
energy densities lower than about 10−14 J m−3 to solve this problem is em-
phasized. Therefore we undertake a search for such GRGs on the southern
sky hemisphere using the SALT. In this paper we present a selected sample
of the GRG candidates and the first deep detections of distant host galaxies,
as well as the low-resolution spectra of the galaxies identified on the DSS
frames. The data collected during the Performance Verification (P -- V) phase
show that 21 of 35 galaxies with the spectroscopic redshift have the pro-
jected linear size greater than 1 Mpc (for H0=71 km s−1Mpc−1). However
their redshifts do not exceed the value of 0.4 and the energy density in only
two of them is less than 10−14 J m−3. A photometric redshift estimate of one
of them (J1420−0545) suggests a linear extent larger than 4.8 Mpc, i.e. a
larger than that of 3C236, the largest GRG known up to now.
Key words: Galaxies:active -- Galaxies:interactions -- Galaxies:
inter-
galactic medium
1. Introduction
In the adiabatically expanding Universe filled with a hot, diffuse and
uniform IGM, the IGM pressure should increase with redshift as pIGM(z) =
p0(1+z)5, where p0 is the present-day pressure (cf. Subrahmanyan & Saripalli
1993). Various methods have been employed in determining the density and
pressure of the IGM. The most effective are X-ray observations of the hot
gas around radio sources of Fanaroff-Riley type II (FRII) yielding direct
estimates of the gas density (Crawford et al. 1999; Hardcastle et al. 2002;
Croston et al. 2004, 2005). Assuming an equation of state of this gas,
a value of pIGM can be specified. However, both the AGN and the large-
scale radio structure likely contribute to the X-ray emission (cf. Kaiser &
Alexander 1999), thus the properties of the environment may be considerably
influenced by the presence of the radio source itself. As the IGM forms
the environment of the radio source, the internal physical properties of its
diffuse lobes were expected to reflect directly pIGM (z). However, the cosmic
background measurements (e.g. with COBE) ruled out an IGM with the
1
pressure indicated by typical radio structures. This suggested that these
regions were not thermally confined. Indeed, the very existence of apparent
hot spots at the edges of double FRII-type radio sources indicates that the
jets ejected from the AGN encounter resistance to their propagation, and
these advancing hot spots are likely confined by ram-pressure of the IGM.
But the more tenuous material, outside the jets, forming a bridge between the
nucleus and the radio lobes is expected to have attained an equilibrium state
in which the energy density of the particles and magnetic field is balanced,
and the pressure of the relativistic plasma equals the pressure of the gaseous
environment. Consequently, a determination of physical conditions in these
diffuse bridges of large-size double radio sources (being away from the ram-
pressure confined the source heads) gives an independent tool of probing the
pressure of the IGM.
This physical conditions, specifically the internal pressure, can be deter-
mined from the dynamical considerations.
In the analytical model of the
source dynamics of Kaiser & Alexander (1997) and its further modifications
(e.g. Blundell et al. 1999) the pressure is dependent on the gas density,
power of the jet, and age of the radio structure. As a certain determination
of these parameters is not easy, the internal pressure is usually estimated
from the minimum energy condition corresponding to an equipartition of en-
ergy between the relativistic particles and the magnetic field. The minimum
energy can be calculated from the total luminosity and the size (volume)
of a given radio source (e.g. Pacholczyk 1970; Miley 1980). The largest
extragalactic radio sources (D>1 Mpc, frequently called Giants or Giant
Radio Galaxies [GRGs] though a small fraction of them are classified as
QSOs), due to their very extended radio lobes and sometime bridges link-
ing the opposite lobes, offer a unique probe the intergalactic and intracluster
environment directly on Mpc scales allowing measurements of the IGM struc-
ture, pressure and their evolution out to high redshifts (cf. Strom & Willis
1980; Subrahmanyan & Saripalli 1993; Mack et al. 1998).
Subrahmanyan & Saripalli (1993) contended that the lower and upper
limit values of the present-day IGM pressure are 0.5 and 2×10−15 N m−2,
respectively, but further investigations raised those values at least twice (e.g.
Ishwara-Chandra & Saikia (1999). The problem of the cosmic pressure evo-
lution was also considered by Schoenmakers et al.
(2000) who found no
evidence for such a strong evolution up to the redshift of about 0.3. They
concluded the apparent pressure increase with redshift can be explained by
two effects: the use of flux-limited samples and the method by which the lobe
pressure is calculated. The latter aspect is discussed more in details in Sec-
tion 6. Therefore, in order to verify the hypothesis about the IGM pressure
evolution concordant with (1+z)5, it is necessary to find GRGs with the lobe
energy densities of < 10−14 J m−3 at redshifts higher than about 0.5. The
project how to do that and the selected sample of GRG candidates are pre-
sented in Section 2. The imaging and spectroscopic observations conducting
either with the SALT or with the 1.9m SAAO telescope are described in Sec-
2
tions 3 and 4. Preliminary physical and geometrical parameters of the sample
sources are determined in Section 5, while the equipartition energy density
in their lobes (proportional to an internal lobe pressure) vs.
the sources'
redshift and morphology of the GRGs vs. properties of the environment are
analyzed in Section 6.
2. High-redshift GRGs and description of the project
Since majority of known GRGs are at rather low redshifts (z≤0.25), it
was long presumed that giants did not exist at significantly higher redshifts,
especially because both the median and maximum source size were suggested
to decrease as fast as (1 + z)−3 (e.g. Gopal-Krishna & Wiita 1987). Such
a result could be explained by systematic (1 + z)3 increase of density of an
uniform IGM confining the source size. Contrary, high-redshift GRGs might
be tracers of relatively low density regions, not fully virialized yet, existing
at early cosmological epochs. Unfortunately, unbiased samples of GRGs are
very difficult to assemble because of surface-brightness selection effects. At
high redshifts, the diffused bridges of emission are likely to be significantly
influenced by the inverse-Compton losses against the cosmic background,
which would affect the appearance and identification of a GRG at these
redshifts.
The first attempt to discover GRGs at (z ≥ 0.3) was made by Cotter et
al. (1996) who used the 7C radio survey to select Giant candidates and found
13 large double radio sources at redshifts up to z ≈0.9, 4 of them with D >1
Mpc. A further few distant GRGs were found by Schoenmakers et al. (1998)
proving that such sources do exist at least out to z ∼1. In Table 1 (below)
we assemble all known Giants with z ≥0.5 and D &1 Mpc (calculated with
the cosmological parameters: H0=71 km s−1Mpc−1, Ωm=0.27, and ΩΛ=0.73.
This is worth emphasizing that only two sources in Table 1 have negative
declinations.
In this project we attempt to enlarge significantly the above sample with
high-redshift GRGs on the southern hemisphere, and provide data sufficient
to enlighten the problem of the IGM pressure evolution with redshift. For
this purpose we selected a large sample of GRG candidates using the existing
1.4 GHz radio surveys: FIRST (Becker et al. 1995) and NVSS (Condon et al.
1998). These surveys, which provide radio maps of the sky made with two
different angular resolutions (5" and 45", respectively) at the same observing
frequency, allow (i) an effective removal of confusing sources, (ii) a reliable
determination of the sources' morphological type, and in many cases (iii)
a determination of the compact core component necessary for the proper
identification of the source with its host optical object.
In order to use a possibility of all-season optical observations with the
SALT, the entire sample consists of two parts (subsamples) located on the
northern and southern Galactic hemispheres:
(NGH) lying between 08h20m < RA(J2000) < 16h00m;
-14o <Dec(J2000)< +13o, and
3
IAU name
Table 1: Known giants with z≥0.5 and D&1 Mpc
Survey z
R
D Id.
PKS
7C
B0437-244
B0654+482
J0750+656
8C
B0821+695
B0854+399 B2
7C
B1058+368
B1127-130
PKS
B1349+647
3C292
B1429+160 MC3
B1602+376
B1636+418
B1834+620 WNB
J1951+706
7C
7C
0.84
0.776
0.747
0.538
0.528
0.750
0.634
0.71
1.005
0.814
0.867
0.519
0.550
[mag]
20.5
21.9
Θ
["]
127
135
220
408
166
150
290
134
167
182 >20
130 >20
220
204
21.3
16.0
21.1
log P1.4
[W Hz−1]
27.17
26.39
26.40
26.25
26.71
26.53
27.26
27.62
26.88
26.59
26.49
26.89
26.07
[kpc]
960 QSO
1002
1606 QSO
2576
1038
1100
2033 QSO
960
1346 QSO
1376
1004
1364
1300
(SGH) lying between 21h00m < RA(J2000) < 03h15m;
-14o <Dec(J2000)< +02o.
Thus, an advance of the project is a uniform distribution of the targets
along hour angle enabling consecutive collection of their images and spectra.
A crude estimate of the source's redshift can be made from the Hubble dia-
gram for GRGs published by Schoenmakers et al. (1998). At z=0.5 it gives
R ≈20 mag (the apparent magnitudes in Table 1 confirm this). On the other
hand, a source with the linear projected size of 1 Mpc at z=0.5 will have an
angular size of about 165". Therefore the GRG candidates were chosen to
fulfill the following selection criteria: they have
(1) the morphological type of FRII,
(2) an angular separation between the brightest regions on the maps Θ >
165",
(3) 1.4-GHz flux density on the NVSS map S >50 mJy in the NGH subsam-
ple, and S >40 mJy in the SGH subsample, and
(4) is optically identified with an object fainter than 20 mag in the R-band.
The above criteria selected rather small number of high-redshift GRG
candidates, especially in the NGH subsample. Therefore in both subsamples
we include also other radio source candidates with host galaxies brighter than
20(R) mag which will be analyzed as comparison closer GRGs. The optical
spectroscopy of these latter sources can be conducted with smaller telescopes.
In Table 2 we list numbers of the GRG candidates in both subsamples. The
observational data on 34 of 39 radio sources in the subsample NGH and
12 of 41 in the subsample SGH, i.e. those for which preliminary physical
parameters could be determined using the data collected during the P-V
phase, are given in Table 3. The consecutive columns of Table 3 give:
Col. 1: IAU name at epoch J2000;
4
Col. 2: Fanaroff-Riley morphological type. "D" indicates a diffuse morphol-
ogy of the radio lobes without any detectable hot spot(s);
Col. 3: 1.4 GHz total flux density in mJy from the NVSS survey;
Col. 4: radio spectral index between 1.4 and 4.9 GHz;
Col. 5: 1.4 GHz flux density of the radio core from the FIRST survey;
Col. 6: largest angular size in arcsec;
Col. 7: optical identification. "G" - galaxy, "Q" - quasar;
Col. 8: its apparent R-band magnitude from the DSS data base;
Col. 9: redshift. A value in parenthesis is the photometric redshift esti-
mate calculated from the log z -- R relation (Hubble diagram) derived from
the spectroscopic redshift and K-corrected magnitudes of the galaxies in Ta-
ble 3. Our relation is consistent with the Hubble diagram for a majority of
known GRGs published by Lara et al. (2001) [their equation (1) and figure
4];
Col. 10: redshift references. (1) our SAAO observations, (2) our SALT ob-
servations, (3) Sloan Digitized Sky Survey [SDSS] Adelman-McCarthy et al.
(2007), (4) Jones et al. (2005), (5) Colless et al. (2001), (6) Machalski et al.
(2001), (7) Bhatnagar et al. (1998), (8) Stickel et al. (1994), (9) Becker et
al. (2001).
Table 2: Content of the subsamples
Subsample Redshift already to be determined to be determined Total
NGH
SGH
available
13
7
with the SALT
16
29
with other tel.
10
5
39
41
5
Table 3: The sample. The entries determined in this paper are marked in
bold face.
LAS Opt. RDSS
[mag]
id.
["]
z
Ref.
to z
Name
FR
type
S1.4
[mJy]
α
Score
[mJy]
NGH:
J0824+0140
J0903+1208
J0922+0919
J0925−0114
J0947−1338
J1005−1315
J1014−0146
J1018−1240
J1021−0236
J1021+1217
J1021+0519
J1048+1108
J1049−1308
J1058+0812
J1101−1053
J1108+0202
J1126−0042
J1130−1320
J1213−0500
J1253−0139
J1328−0129
J1328−0307
J1334−1009
J1354−0705
J1411+0619
J1420−0545
J1445+0932
J1445−0540
J1457−0613
J1459−0432
J1520−0546
J1528+0544
J1540−0127
J1543−0112
SGH:
J2234−0224
J2239−0133
J2345−0449
J0010−1108
J0037+0027
J0039−1300
J0042−0613
J0129−0758
J0134−0107
J0202−0939
J0300−0728
J0313−0632
IID
II
IID
IID
II
IID
II
I/II
II
II
II
IID
IID
II
II
II
II
II
IID
II
II
II
II
II
II
II
II
II
II
IID
II
II
II
II
II
II
IID
IID
II
II
II
I/II
I/II
II
II
II
52
113
104
71
977
58
225
276
218
138
135
235
193
49
92
209
166
1136
>38
171
343
201
1893
95
115
89
144
235
507
62
63
247
209
117
73
134
162
55
134
197
1204
102
287
431
51
165
0.85
0.76
1.27
0.99
1.02
0.89
1.03
0.61
1.2
0.67
0.77
0.95
1.05
1.06
0.81
1.08
0.70
0.91
?
0.87
1.0
290 G
320 G
434 G
832 G
1514 G
260 G
292 G
570 G
357 G
865 G
835 G
294 G
420 G
367 G
852 G
556 G
239 G
297 Q
535 G
231 G
323 G
810 G
850 G
240 G
310 G
1045 G
315 G
420 G
253 G
405 G
1320 G
824 G
295 G
175 G
198 Q
1225 G
1025 G
550 G
299 G
640 G
390 G
720 G
820 G
154 G
300 G
195 Q
5.9
<1.
1.3
12.8
20.7
20.4
24.8
7.6
16.9
5.8
3.2
?
<1.
<6.
?
1.7
18.2
14.3
<1.
16.6
8.8
100.
2.2
22.0
2.7
6.4
12.6
0.73
10.7
20.3
10.3
0.5
6.1
1.8
3.5
5.4
2.1
0.4
1.2
6
16.91
18.73
18.11
12.30
12.70
19.34
16.54
14.66
18.50
15.49
16.26
15.71
16.05
19.74
16.19
15.57
18.10
15.98
15.69
18.33
16.91
16.82
13.58
17.30
20.00
19.65
14.89
19.44
16.64
18.76
12.35
11.5
16.57
20.78
18.63
13.45
13.4
12.57
20.55
12.93
16.47
14.0
12.25
20.2
20.44
19.20
0.2125
(0.27)
(0.23)
0.0732
0.0800
(0.31)
0.1986
0.0777
0.2917
0.1294
0.1562
0.1570
(0.15)
(0.33)
(0.15)
0.1574
0.3317
0.6337
0.0857
(0.24)
0.1513
0.0860
0.0838
(0.19)
0.3591
(0.32)
0.094
0.3666
0.1671
(0.27)
0.0607
0.040
0.1490
0.3680
0.55
0.0865
0.0757
0.0773
(0.42)
0.1060
0.123
0.0991
0.0790
(0.37)
0.4905
0.389
1
4
1
5
6
2
3
3
3
3
2
7
1
3
3,6
4,8
2
1
2
6
1
1
3
2
9
1
4
3
1
1
4
3
3
3
3. Optical imaging and identification
3.1. DSS identification
To identify the GRG candidates whose radio core is known with a host
optical object, first at all we used the DSS, i.e.
the digitized POSS and
UKST surveys. For all NGH sources with cores, we have identified their cores
with galaxies; only one sample source is known radio quasar (J1130−1320).
For the remaining NGH candidates -- the galaxies given in Table 3, being
the only objects in a middle of the radio structure, thus are very probable
identifications. Quite different situation is in the SGH subsample whose sky
area is almost not covered by the FIRST survey. Therefore for a certain
optical identifications high-resolution radio observations are necessary. The
SGH sources, where at the position of detected radio core no optical object
had been visible in the DSS data base, were scheduled for a deep detection
with the SALT.
3.2. Preliminary deep imaging and photometry with the SALT
During the Performance Verification (P-V) phase two blank fields in the
SGH subsample were observed using the SALTICAM. Series of 60sec-long
exposures were taken through the B, V and R filters. The frames were cor-
rected for overscan, cross-talk and mosaiced. Because flat-field frames were
not available for the tracker position close to the objects' position, the sur-
face functions were fitted and then subtracted from the frames in order to
obtain possibly flat and uniform background. Instrumental magnitudes of
objects in the frames were determined using the aperture method. Apparent
magnitudes of these objects were estimated using magnitudes of stars in each
field for which their standard magnitudes were available from the APM cat-
alogue (http://www.ast.cam.ac.uk/∼apmcat/). The astrometric calibration
was done transforming the instrumental coordinates of these stars in a given
frame into their sky coordinates in the DSS data base. These coordinates
were not corrected for the proper motion and parallax.
The deep optical images of the field around the sources J0037+0027 and
J0202−0939 are shown in Figs. 1a, b. Overlaid are intensity-contours of the
radio maps made with the combined FITS data of the FIRST and NVSS
surveys. On both images a faint galaxy coincides with the radio core. The
resulting magnitudes and their errors are given in Table 4.
3.3. Auxiliary photometry of J1420−0545
Because of a suspicion that the radio galaxy J1420−0545 can be the
largest GRG observed up to now (cf. Introduction), we made an auxiliary
optical photometry for this galaxy. The observations were made using the
60cm telescope of the Pedagogical Academy at Mt. Suhora. The resulting
magnitudes in the UBV RI photometric system, given in Table 4, are cor-
rected for the atmospheric extinction only. As the galactic latitude of this
galaxy is 50 degr, the foreground Galactic extinction is close to zero (Sandage
1972). In order to be consistent with the DSS data, these magnitudes are not
7
K-corrected. Our R-band magnitude is thus identical as the corresponding
value in Table 3 taken from the DSS data base. The optical spectral index of
about 4.0 resulting from our VRI magnitudes also suggests a redshift higher
than 0.3 for any of the evolutionary models for early-type galaxies (Bruzual
1983).
J0202-0939
)
0
0
0
2
J
(
I
N
O
T
A
N
L
C
E
D
I
J0037+0027
00 29 30
00
28 30
00
27 30
00
26 30
00
25 30
00 38 02
00
-09 37 30
a)
38 00
30
)
0
0
0
2
J
(
I
N
O
T
A
N
L
C
E
D
I
39 00
30
40 00
30
37 58
52
RIGHT ASCENSION (J2000)
56
54
50
48
02 02 28
27
26
25
22
RIGHT ASCENSION (J2000)
24
23
b)
21
20
Figure 1: Deep images of the optical fields taken with the SALTICAM: a)
around the source J0039+0027 and b) around the source J0202−0939. The
solid contours indicate the 1.4 GHz brightness distribution observed in the
NVSS + FIRST surveys
4. Optical spectroscopy
4.1. Low-resolution spectroscopy with the SALT
In order to determine physical parameters of a given sample source, e.g.
its radio and optical luminosity, linear size and volume, energy density, etc., a
distance to the host galaxy (or QSO) must be known which can be calculated
from its redshift. Therefore, using the low-resolution spectrograph (PFIS)
during the P-V observations we attempted to detect optical spectrum and
determine redshifts of 6 faint host galaxies. The observations were made
in series between May 30th 2006 and June 24th 2006. The spectra were
obtained with the PFIS in Grating Spectroscopy mode using the standard
8
I[mag]
R[mag]
Source
J0037+0027
J0202−0939
J1420−0545
Table 4: IRV B magnitudes of the detected galaxies
B[mag]
V [mag]
--
--
20.55±0.16
--
18.44±0.08
19.65±0.08
21.69±0.15
21.36±0.15
(21.0±0.2)
23.12±0.13
22.89±0.12
--
surface-relief grating SR300, slit width of 1.5 arcsec, granting angle of 7
deg, and the binning of 2×2. These settings resulted in the typical useful
spectral range about 3800A -- 8500A and the dispersion of 3.05A per pixel.
Because of the certain space between the CCD chips there is a gap between
6280A and 6440A in all the SALT spectra. The pre-reduced (i.e. bias and
overscan subtracted, trimmed, cross-talk corrected, mosaiced and cosmic-ray
removed) files were further reduced using the standard IRAF tasks. All the
spectra were wavelength calibrated using the CuAr arc lamp exposures and
one-dimensional spectra were extracted. Gaussian profiles were fitted to the
lines visible in the spectra and the wavelengths corresponding to the profiles'
centers were used to calculate the redshift.
4.2. Low-resolution spectroscopy with the 1.9m SAAO telescope
9 host galaxies brighter than about 16 mag (R) were observed with the
1.9m telescope equipped with the Cassegrain Grating Spectrograph in the
long-slit mode. The galaxies were observed with low-resolution grating 8,
slit width of 300µm (2.5 arcsec) and the grating angle of 14.7 deg resulting
in the spectral range 5300A -- 7600A and the dispersion of 1.66A per pixel.
Exposures of the CuAr arc lamp were made every 30 min. Every night
selected spectrophotometric standards were also observed.
Reduction of the spectra were carried out using the standard IRAF longslit
package. All the spectra were corrected for bias, flat-field and cosmic-rays, as
well as wavelength and flux-density calibrated. In order to obtain the result-
ing spectra with a better S/N ratio, all calibrated one-dimensional spectra
of a given galaxy were combined.
4.3. The resulting spectra
The final, low-resolution, 1 -- D spectra of 14 observed galaxies are shown
in Figs. 2a -- d. The emission lines and/or absorption bands detected, as
well as the resultant redshift of these features are listed in Table 5. Because
the spectra taken with the SALT were not flux-density calibrated, relevant
ordinates in Figs. 2 give the normalized counts only. Most of the flux-
density calibrated spectra are typical of elliptical galaxies whose continuum
emission with the broad absorption bands is dominated by evolved giant
stars. The emission lines mostly detected are [OII]λ3727 and [OIII]λ4960
and λ5007. Other forbidden lines present in a few spectra are [NII]λ6585, as
well as [NII]λ6718 and λ6733. Also weak Balmer lines are detected in five of
the spectra, however the Hα line redshifted by more than 1.25 was usually
9
beyond the wavelength range observed.
5. Physical parameters
In order to analyse and discuss our preliminary results on the IGM evo-
lution, we determine or estimate some global physical and geometrical pa-
rameters for the sample sources. These parameters are defined and given
in Table 6. The entries in parentheses are the approximated values for the
sources with the photometric redshift estimate given in Table 3. The seven
columns of Table 6 give:
Col. 1: Repeated IAU name;
2: Logarithm of radio luminosity at the emitted frequency of 1.4
Col.
GHz calculated with the cosmological parameters: H0=71 km s−1Mpc−1,
Ωm=0.27, ΩΛ=0.73;
Col. 3: Projected linear size;
Col. 4: Average magnetic field calculated from the revised equipartition and
minimum energy formula of Beck & Krause (2005) under assumption of a
cylindrical geometry of the extended radio emission with the base diameter
set equal to the average width of the lobes and K0≡ (Ep/Ee)α0=52 for the
injection spectral index of α0=0.525;
Col. 5: Equipartition energy density under the same assumption;
Col. 6: Arm separation ratio: the separation of the farther hotspot or the
lobe's boundary from the core or the identified optical galaxy to the nearer
one;
Col. 7: Core prominence defined as the ratio fc = Pcore/(Ptotal − Pcore),
where both Pcore and Ptotal are determined at the emitted frequency of 1.4
GHz using the data given in Table 3, however for the sample sources without
a certain spectral index the value of 0.9 has been assumed, and the relevant
value of 0.0 has been taken for the cores.
An estimation of the magnetic field strength in extended lobes of dou-
ble radio sources still is a controversial matter.
In majority of previously
published papers the classical energy equipartition approach of Pacholczyk
(1970) and its formalism of Miley (1980) was used for this purpose. In the
Pacholczyk's formula the equipartition magnetic field strongly depends on
the hardly known ratio of energy in heavy particles to that in the electrons
(+positrons) k; in most of the above papers k=1 was assumed and the radio
spectrum was integrated between a defined frequency range, e.g. between 10
MHz and 100 GHz. In a slightly different approach to the equipartition field
calculations, the spectrum is integrated between frequencies corresponding
to a minimum and a maximum Lorentz factor for the relativistic electrons
(cf. Croston et al. 2005). To overcome the problem of uncertain value of k,
Beck & Krause (2005) have proposed a revised formula for the magnetic field
strength applicable in the simple case that the number density ratio K0 of
protons and electrons is constant which is valid in a limited range of particle
energies. They estimate that the value of this parameter is 40 < K0 < 100
depending on a value of the injection spectral index α0. The above three
10
different estimates of average magnetic field in the lobes of selected GRGs
are compared between themselves in Konar et al. (2007).
6. Discussion of the results
6.1. Correlation between energy density (pressure) in the lobes and redshift
Since an average internal lobe pressure is proportional to the energy den-
sity, in Fig. 3 we plot logarithm of the latter vs. the source's redshift (1+z)
for the sample of our GRGs, and supplement it with the corresponding data
for two other samples: a limited sample of already known GRGs as well as
a sample of "normal-sized" FRII-type radio galaxies taken from Machalski
& Jamrozy (2006). Because the equipartition magnetic field calculated us-
ing the Beck & Krause formula is larger than the Miley's value by a factor
of ∼3, the resulting energy densities are one order larger than the classical
values. Therefore these values in Fig. 3 (and Table 6) are about 10 times
higher than corresponding values calculated with the Miley's formula and
the lowest values of ueq are about 10−14 J m−3 instead of about 10−15 J m−3
like in Subrahmanyan & Saripalli (1993) or Cotter (1998). However this dif-
ference can be neglected if we compare the sources with similarly calculated
magnetic fields.
This is clearly seen in Fig. 3 that we find 13 GRGs with the values of
ueq lower than the lowest values found up to now, though most of them
characterize the low-redshift sources at z < 0.2. An exception is the source
J1420−0545. The solid bar indicates an uncertainty of its redshift corre-
sponding to the standard deviation of about 1 mag in the distribution of
absolute optical magnitudes of galaxies hosting the GRGs (cf. Lara et al.,
2001). The photometric redshift estimate of 0.32 implies a projected linear
size of over 4.8 Mpc which, if real, would be larger than the largest GRG
known up to now, i.e. 3C236! Unfortunately, our attempts to obtain its op-
tical spectrum either with the SALT or to get observing time at other large
telescopes were unsuccessful. The dashed straight line marks the ueq ∝(1+z)5
relation.
Anyhow the data presented in this paper confirm the known strong in-
crease of the equipartition energy density with increasing redshift (cf. Sub-
rahmanyan & Saripalli 1993; Cotter 1998). To investigate how the observed
ueq vs. (1+z) correlation reflects an expected change in the environment of
GRGs one has to check how the selection effects affect the diagram shown in
Fig. 3. According to the classical formula, the equipartition energy density is
related to the total power and the volume of a radio source by ueq ∝ (P/V )4/7.
Since all the sources plotted in the diagram belong to the flux-density limited
samples, therefore sources at higher redshift have, on average, higher radio
powers and thus higher energy density at equal source volumes. Moreover,
we can expect that an average volume of sources decreases with redshift as
does their average linear size. Following Schoenmakers et al. (2000) we cal-
culate the expected ueq as a function of (1+z) for sources with flux density
of 45 mJy, i.e. a mean value of the limiting fluxes in the NGH and SGH
11
subsamples, and a volume corresponding to cylinder length of 1 Mpc and its
diameter of 167 kpc. The resulting relation is shown with the solid curve
whose shape simply indicates the log P -- log(1+z) relation for a constant flux
density and the radio spectral index of unity.
In our previous paper (Machalski & Jamrozy 2006) we showed that per-
forming a statistical test for correlations between two variables in the pres-
ence of one or two other variables one can examine whether a residual cor-
relation between the above two variables (parameters) is still present when
the third (or the third and fourth) is (are) held constant. In particular, we
showed that the equipartition energy density calculated for the sample radio
sources, but transformed to a reference total power and a reference linear
size, practically did not correlate with redshift. It seems that supplementing
their data (crosses and full dots in Fig. 3) with further GRGs investigated in
this paper (open circles in Fig. 3) do not change the situation significantly.
6.2. Morphology of GRGs vs. the environment
Projected linear sizes of the largest FRII-type radio sources, considerably
exceeding those predicted (at different redshifts) by the very early dynamical
model of Gopal-Krishna & Wiita (1987), already suggested that the IGM may
be inhomogeneous and large-scale density fluctuations, especially at high
redshifts, can play a significant role in the time evolution of those sources.
The unified scheme for radio sources with active galactic nuclei (AGN)
(e.g. Barthel 1989; Gopal-Krishna, Kulkarni & Mangalan 1994; Urry &
Padovani, 1995) predicts how the appearance of sources evolving in a ho-
mogeneous environment depends on orientation of the jets' axis towards the
observer's line of sight. This appearance includes the arm separation ratio
Rθ, the core prominence parameter, fc (both defined in Section 5), and the
flux density ratio in radiation from the opposite lobes (not analysed in this
paper). Within this scheme, the effects of relativistic beaming cause that the
sources at small inclination angles should have more prominent cores and
higher asymmetry ratios in the arm length and flux density compared with
those at larger angles.
The evident lack of a statistical correlation between Rθ and fc in our sam-
ple (the correlation coefficient of −0.38 suggesting even an anticorrelation)
strongly supports a common conviction that the inclination angle of majority
of GRGs is close to 90o. Therefore, because the lobes of GRGs lie well beyond
the boundary of the parent optical galaxies, their symmetry/asymmetry pa-
rameters may reflect eventual environmental heterogeneity. Ishwara-Chandra
& Saikia (1999) and Schoenmakers et al. (2000) noted that the GRGs tended
to be marginally more asymmetric than smaller sources of a comparable radio
luminosity. For example, Ishwara-Chandra & Saikia noted that the median
value of the arm separation ratio in their sample of GRGs is about 1.39
compared with about 1.19 for 3CR galaxies of similar luminosity but smaller
sizes. According to the unified scheme, the expected arm separation ratio
for a source inclined at i > 45o to the line of sight is only about 1.15 for the
12
jet head advance speed of ∼0.1 (cf. Scheuer 1995). As the inclination angles
of GRGs very likely are close to 90o, the arm asymmetry ratio (frequently
much higher than one) should be related to the environmental properties
rather than to an orientation towards the observer.
The radio data available for the GRGs investigated in this paper allow
us to study a distribution of Rθ, and (within a limited range) fc. The
median value of 1.25±0.03 in the distribution of Rθ in our sample sup-
ports the Ishwara-Chandra & Saikia's conclusion that the apparent arm
asymmetry in the GRGs "could possibly be caused by interaction of the
energy-carrying beams with cluster-sized density gradients far from the par-
ent galaxy". The presented analysis of the sample of faint low-luminosity
GRGs whose lobes lie at least at 0.5 -- 1 Mpc from the parent AGN strongly
suggests that these sources, after probing a denser central-halo environ-
ment, might further evolve in exceptionally low-density intergalactic medium
to some degree being "voids" in the hot IGM. Therefore, search for low-
luminosity, high-redshift GRGs and the analysis of their physical and struc-
tural parameters, which is the aim of our SALT project, should help to
study these density fluctuations and find their relation to the theoretically
predicted cosmological evolution of the IGM density due to the expansion of
the Universe.
Bearing in mind the above we can conclude as follows:
-- If the internal pressure in diffuse lobes really counterbalances the IGM
pressure, the data collected till now do not provide a certain observational
evidence for the strong cosmological evolution of the latter in the form
pIGM(z) ∝ (1 + z)5. However, if the lobes of the most extended and dif-
fused sources are still overpressured in respect to the IGM, especially at low
redshifts, any conclusion about the properties of the IGM will not be reliable.
-- Deduced low internal pressure in the lobes of GRGs, their observed
morphology and the arm asymmetry strongly suggest an existence of deep
gradiants in the IGM density with the scale of a few Mpc. However, because
an internal pressure in the lobes is proportional to their avarage energy den-
sity, a value of this pressure depends of how the magnetic field strength is
determined. On the basis of the Chandra and XXM-Newton data on X-ray
emission from the lobes of classical double radio sources, Croston et al. (2005)
found that "the measured X-ray flux can be attributed to inverse-Compton
scattering of the cosmic microwave background radiation with magnetic field
strengths in the lobes between 0.3Beq and 1.3Beq, where the value Beq corre-
sponds to equipartition between the electrons and magnetic field, assuming
a filling factor of unity". Therefore the equipartition magnetic field, energy
density, and thus internal lobe-pressure estimates, at least in "normal-sized"
sources are likely close to their true values. Unfortunately, X-ray data are
not yet available for the lobes of GRGs, in particular for those of low radio
luminosity.
-- There is still a hope that detection of the most distant and low-luminosity
13
GRGs, which is the aim of our SALT project, can help in an observational
confinement on the hypothesis of cosmological evolution of the IGM. The
most promising is our SGH subsample consisting of many GRG candidates
optically not identified on the DSS frames, thus with apparent magnitudes
of R > 20.5 mag.
Acknowledgements The authors thank Dr Alexei Kniazev and the remain-
ing SALT staff for their effort in making the PFIS observations. D. K.-W.
acknowledges the hospitality of the SAAO authorities and Dr Stephen Pot-
ter for helping in the observations. We also also thank Micha l Siwak for
the Mt. Suhora observations. This project was supported by MNiSW with
funding for scientific research in the years 2005 -- 2007 under contract No.
0425/PO3/2005/29. Support from SALT International Network grant No.
76/E-60/SPB/MSN/P-03/DWM 35/2005-2007 is also acknowledged.
References
Adelman-McCarthy, J.K., Agueros, M.A., Allam, S.S., et al., 2007, ApJS,
172, 634
Barthel, P.D., 1989, ApJ, 336, 606
Beck, R., and Krause, M., 2005, Astron. Nachr., 326, 414
Becker, R.H., White, R.L., and Helfand, D.J., 1995, ApJ, 450, 559
Becker, R.H., White, R.L., Gregg, M.D., et al., 2001, ApJS, 135, 227
Bhatnagar, S., Gopal-Krishna, and Wisotzki, L., 1998, MNRAS, 299, L25
Blundell, K.M., Rawlings, S., and Willott, C.J., 1999, AJ, 117, 766
Bruzual, G.A., 1983, Rev. Mexicana A&A, 8, 63
Colless, M.M., Dalton, G.B., Maddox, S.J., et al., 2001, MNRAS, 328, 1039
Condon, J.J., Cotton, W.D., Greisen, E.W., et al., 1998, AJ, 115, 1693
Cotter, G., 1998, Astrophysics and space science library (ASSL), Eds. M.N.
Bremer et al. (Dordrecht: Kluwer Academic Publishers), p. 233
Cotter, G., Rawlings, S., and Saunders, R., 1996, MNRAS, 281, 1081
Croston, J.H., Birkinshaw, M., Hardcastle, M.J., and Worrall, D.M., 2004,
MNRAS, 353, 879
Croston, J.H., Hardcastle, M.J., Harris, D.E., et al., 2005, ApJ, 626, 733
Crawford, C.S., et al., 1999, MNRAS, 308, 1195
Gopal-Krishna, Kulkarni, V.K., and Mangalam, A.V., 1994, MNRAS, 268,
459
Gopal-Krishna, and Wiita, P.J., 1987, MNRAS, 226, 531
Hardcastle, M.J., Birkinshaw, M., Cameron, R.A., et al., 2002, ApJ, 581,
948
Ishwara-Chandra, C.H., and Saikia, D.J., 1999, MNRAS, 309, 100
Jones, D.H., Saunders, W., Read, M., and Colless, M., 2005, PASA, 22, 277
Kaiser, C.R., and Alexander, P., 1997, MNRAS, 286, 215
Kaiser, C.R., and Alexander, P., 1999, MNRAS, 305, 707
Konar, C., Jamrozy, M., Saikia, D.J., and Machalski, J., 2007, MNRAS, (in
14
print, astro-ph/0709.4470)
Lara, L., M´arquez, I., Cotton, W.D., Feretti, L., et al., A&A, 378, 826
Machalski, J., and Jamrozy, M., 2006, A&A, 454, 95
Machalski, J., Jamrozy, M., and Zola, S., 2001, A&A, 371, 445
Mack, K.H., Klein, U., O'Dea, C.P., et al., 1998, A&A, 329, 431
Miley, G.K., 1980, ARA&A, 18, 165
Pacholczyk, A.G., 1970, Radio Astrophysics, Freeman and Co., San Fran-
cisco
Sandage, A., 1972, ApJ, 178, 25
Scheuer, P.A.G., 1995, MNRAS, 277, 331
Schoenmakers, A.P., van der Laan, H., Rottgering, H.J.A., and de Bruyn,
A.G., 1998 ASP Conf. Ser. No. 146, Eds. S. D'Odorico A., Fontana,
and E. Giallongo, p.84
Schoenmakers, A.P., Mack, K.-H., de Bruyn, A.G., et al., 2000, AA&S, 146,
293
Stickel, M., Meisenheimer, K., and Kuhr, H., 1994, A&AS, 105, 211
Strom, R.G. and Willis, A.G., 1980, A&A, 85, 36
Subrahmanyan, R. and Saripalli, L., 1993, MNRAS, 260, 908
Urry, C.M. and Padovani, P., 1995, PASP, 107, 803
15
Line/band
detected
[OII]3727
CaII 3935
CaII 3970
G band
Mg band
NaD band
λobs
[A]
5065.9
5350.7
5396.0
5853.8
7032.6
8008.2
5661.5
6450.8
J1445+0932 Mg band
NaD band
J1445−0540
[OII]3727
[OIII]4959
[OIII]5007
5094.1
6777.2
6842.3
J1520−0546 Mg band
NaD band
[NII]6585
[SII]6733
5489.7
6252.5
6986.5
7142.5
z
0.3592
0.3598
0.3592
0.3598
0.3584
0.3582
0.3591±0.0008
0.0936
0.0941
0.094±0.0010
0.3668
0.3664
0.3665
0.3666±0.0003
0.0604
0.0605
0.0610
0.0608
0.0607±0.0003
J1528+0544 Mg band
NaD band
5381.4
6135.4
0.0395
0.0406
0.040±0.0010
J1543−0112
[OII]3727
[OIII]4364
[OIII]4960
[OIII]5007
5100.4
5971.8
6784.4
6848.8
0.3685
0.3684
0.3678
0.3678
0.3681±0.0005
J2239−0133 Mg band
NaD band
Hα
5622.9
6406.5
7157.3
0.0861
0.0866
0.0865
0.0865±0.0006
Table 5: Lines detected and redshift of the galaxies
z
Source
J1411+0619
Source
Line/band
detected
J0039−1300 Mg band
NaD band
[NII]6550
[NII]6585
[SII]6718
[SII]6733
J0042−0613 Mg band
NaD band
J0824+0140 Hβ
[OIII]4959
[OIII]5007
J0947−1338 Mg band
NaD band
[NII]6550
Hα
[NII]6585
[SII]6718
λobs
[A]
5725.9
6519.9
7244.0
7284.0
7432.0
7446.0
5810.5
6623.5
5894.2
6014.0
6070.6
5591.1
6367.0
7072.9
7089.8
7112.2
7255.4
J1021−0236
[OII]3727
[OIII]5007
4814.9
6466.5
J1126−0042
[OII]3727
[OIII]4960
[OIII]5007
Hα
4965.1
6602.0
6665.7
8742.9
J1213−0500 Mg band
NaD band
Hα
[NII]6585
5618.7
6399.0
7127.4
7153.6
0.1060
0.1058
0.1060
0.1061
0.1062
0.1059
0.1060±0.0002
0.1224
0.1234
0.123±0.0010
0.2125
0.2125
0.2124
0.2125±0.0001
0.0800
0.0799
0.0798
0.0803
0.0801
0.0800
0.0800±0.0002
0.2919
0.2915
0.2917±0.0005
0.3322
0.3310
0.3313
0.3321
0.3317±0.0007
0.0853
0.0853
0.0860
0.0863
0.0857±0.0005
16
1e-15
9e-16
8e-16
7e-16
6e-16
5e-16
4e-16
3e-16
2e-16
1.4e-15
1.2e-15
1e-15
8e-16
6e-16
4e-16
2e-16
0
-2e-16
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
1
7
e
a
c
h
p
a
n
e
l
J
0
0
4
2
−
0
6
1
3
,
J
0
8
2
4
+
0
1
4
0
,
J
0
9
4
7
−
1
3
3
8
,
w
h
o
s
e
I
A
U
n
a
m
e
s
a
r
e
m
a
r
k
e
d
i
n
F
i
g
u
r
e
2
a
:
O
p
t
i
c
a
l
l
o
w
-
r
e
s
o
l
u
t
i
o
n
s
p
e
c
t
r
a
o
f
t
h
e
s
o
u
r
c
e
s
:
J
0
0
3
9
−
1
3
0
0
,
]
5
8
5
6
[
I
I
N
]
0
5
5
6
[
I
I
N
]
3
3
7
6
[
I
I
S
+
]
8
1
7
6
[
I
I
S
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
4.5e-16
4e-16
3.5e-16
3e-16
2.5e-16
2e-16
1.5e-16
1e-16
8e-16
7e-16
6e-16
5e-16
4e-16
3e-16
2e-16
J0042-0613
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
5500
6000
6500
Wavelength [A]
7000
7500
J0947-1338
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
]
5
8
5
6
[
I
I
N
α
Η
]
3
3
7
6
[
I
I
S
+
]
8
1
7
6
[
I
I
S
5000
5500
6000
6500
Wavelength [A]
7000
7500
1.1e-15
J0039-1300
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
5500
6000
6500
7000
7500
Wavelength [A]
J0824+0140
]
7
0
0
5
[
I
I
I
O
]
0
6
9
4
[
I
I
I
O
β
Η
5000
5500
6000
6500
7000
7500
Wavelength [A]
10
J1021-0236
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
8
6
4
2
0
-2
-4
-6
]
7
2
7
3
[
I
I
O
]
7
0
0
5
[
I
I
I
O
4000
4500
5000
5500
6000
6500
Wavelength [A]
7000
7500
8000
8500
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
6e-16
5e-16
4e-16
3e-16
2e-16
1e-16
0
J1213-0500
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
]
3
3
7
6
[
I
I
S
+
]
8
1
7
6
[
I
I
S
]
5
8
5
6
[
I
I
N
α
Η
5000
5500
6000
6500
Wavelength [A]
7000
7500
1
8
J
1
1
2
6
−
0
0
4
2
,
J
1
2
1
3
−
0
5
0
0
,
J
1
4
1
1
+
0
6
1
9
F
i
g
u
r
e
2
b
:
T
h
e
s
a
m
e
a
s
i
n
F
i
g
.
2
a
b
u
t
f
o
r
t
h
e
s
o
u
r
c
e
s
:
J
1
0
2
1
−
0
2
3
6
,
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
8
7
6
5
4
3
2
1
0
-1
4
3
2
1
0
-1
J1126-0042
]
7
2
7
3
[
I
I
O
]
7
0
0
5
[
I
I
I
O
]
0
6
9
4
[
I
I
I
O
]
3
6
5
6
[
α
Η
4000
5000
6000
Wavelength [A]
7000
8000
J1411+0619
]
7
2
7
3
[
I
I
O
]
]
5
3
9
3
0
7
9
3
[
[
K
H
]
6
0
3
4
[
G
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
4500
5000
5500
6000
6500
7000
7500
8000
8500
Wavelength [A]
9e-16
8e-16
7e-16
6e-16
5e-16
4e-16
3e-16
2e-16
1e-16
0
1e-15
9e-16
8e-16
7e-16
6e-16
5e-16
4e-16
3e-16
2e-16
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
1
9
J
1
4
4
5
−
0
5
4
0
,
J
1
5
2
0
−
0
5
4
6
,
J
1
5
2
8
+
0
5
4
4
F
i
g
u
r
e
2
c
:
T
h
e
s
a
m
e
a
s
i
n
F
i
g
.
2
a
b
u
t
f
o
r
t
h
e
s
o
u
r
c
e
s
:
J
1
4
4
5
+
0
9
3
2
,
J1445+0932
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
5000
5500
6000
6500
Wavelength [A]
7000
7500
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
12
10
8
6
4
2
0
-2
-4
-6
J1520-0546
]
5
8
5
6
[
Ι
Ι
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
Ν
+
α
Η
]
3
3
7
6
+
8
1
7
6
[
I
I
S
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
5000
5500
6000
6500
Wavelength [A]
7000
7500
1.2e-15
1.1e-15
1e-15
9e-16
8e-16
7e-16
6e-16
5e-16
4e-16
3e-16
]
7
2
7
3
[
I
I
O
J1445-0540
]
]
0
6
9
4
7
0
0
5
[
I
I
I
[
I
I
I
O
O
4500
5000
5500
6000
6500
7000
7500
8000
8500
Wavelength [A]
J1528+0544
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
5000
5500
6000
Wavelength [A]
6500
7000
7500
]
1
-
A
1
-
s
2
-
m
c
g
r
e
[
x
u
F
l
7e-16
6.5e-16
6e-16
5.5e-16
5e-16
4.5e-16
4e-16
3.5e-16
3e-16
2.5e-16
J2239-0133
]
7
7
1
5
[
g
M
]
6
9
8
5
[
a
N
α
Η
5500
6000
6500
7000
7500
Wavelength [A]
J1543-0112
]
7
0
0
5
[
I
I
I
O
]
0
6
9
4
[
I
I
I
O
]
7
2
7
3
[
I
I
O
]
4
6
3
4
[
I
I
I
O
4000
4500
5000
5500
6000
6500
7000
7500
8000
Wavelength [A]
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
3.5
3
2.5
2
1.5
1
0.5
0
-0.5
J
2
2
3
9
−
0
1
3
3
F
i
g
u
r
e
2
d
:
T
h
e
s
a
m
e
a
s
i
n
F
i
g
.
2
a
b
u
t
f
o
r
t
h
e
s
o
u
r
c
e
s
:
J
1
5
4
3
−
0
1
1
2
,
2
0
Table 6: Physical parameters. The entries in parentheses concern the sample
sources with the photometric redshift estimate
Name
NGH:
J0824+0140
J0903+1208
J0922+0919
J0925−0114
J0947−1338
J1005−1315
J1014−0146
J1018−1240
J1021−0236
J1021+1217
J1021+0519
J1048+1108
J1049−1308
J1058+0812
J1101−1053
J1108+0202
J1126−0042
J1130−1320
J1213−0500
J1253−0139
J1328−0129
J1328−0307
J1334−1009
J1354−0705
J1411+0619
J1420−0545
J1445+0932
J1445−0540
J1457−0613
J1459−0432
J1520−0546
J1528+0544
J1540−0127
J1543−0112
SGH:
J2234−0224
J2239−0133
J2345−0449
J0010−1108
J0037+0027
J0039−1300
J0042−0613
J0129−0758
J0134−0107
J0202−0939
J0300−0728
J0313−0632
logP1.4
[W Hz−1]
24.83
(25.40)
(25.23)
23.95
25.17
(25.25)
25.38
24.60
25.76
24.77
24.93
25.19
(25.05)
(25.24)
(24.74)
25.13
25.76
27.29
23.82
(25.43)
25.31
24.56
25.50
(25.00)
25.67
(25.44)
24.49
26.02
25.58
(25.14)
23.73
23.95
25.09
25.70
25.94
24.39
24.35
23.90
(25.90)
24.73
25.64
24.39
24.63
(26.28)
25.67
25.93
D
[kpc]
994
(1310)
(1580)
1144
2258
(1175)
949
828
1547
1974
2234
790
(1085)
(1730)
(2200)
1497
1132
2033
849
(868)
841
1290
1322
(750)
1547
(4830)
543
2124
715
(1660)
1525
643
759
887
1266
1960
1454
795
(1650)
1228
852
1300
1209
(785)
1806
1024
Beq
[nT]
ueq×10−14
[J m−3]
Rθ
fc
4.67
(5.74)
(3.52)
1.55
2.47
(5.33)
10.95
6.45
6.61
1.66
1.75
13.27
(6.35)
(2.62)
(1.24)
4.23
10.76
16.07
1.92
(13.27)
11.52
3.11
9.64
(9.82)
5.00
(0.83)
11.70
4.76
24.65
(3.25)
0.77
4.56
12.30
14.37
7.24
1.31
2.32
2.79
(5.85)
4.00
8.92
2.51
2.56
(38.0)
3.12
14.10
1.19
1.35
1.52
1.31
1.13
1.30
1.08
1.42
1.40
1.12
1.56
1.30
1.40
1.25
1.24
1.22
1.77
1.04
1.02
1.25
1.00
1.16
1.11
1.60
1.20
1.08
1.05
1.24
1.30
1.42
1.06
1.39
?
1.30
1.45
2.22
1.22
1.35
1.25
1.14
?
1.10
?
1.41
1.43
1.55
0.105
--
0.010
?
0.020
--
0.086
0.089
0.028
0.124
0.039
0.012
--
--
--
--
0.008
--
--
--
0.045
0.041
0.053
0.020
?
0.024
0.043
0.042
--
0.010
0.192
0.086
0.044
0.003
0.060
0.013
0.021
0.100
0.011
--
--
--
--
0.001
0.017
--
0.224
(0.249)
(0.195)
0.129
0.163
(0.240)
0.344
0.264
0.267
0.134
0.137
0.378
(0.260)
(0.168)
(0.115)
0.213
0.340
0.419
0.144
(0.378)
0.352
0.183
0.322
(0.325)
0.232
(0.095)
0.355
0.227
0.515
(0.187)
0.091
0.222
0.363
0.393
0.279
0.119
0.158
0.173
(0.250)
0.208
0.310
0.164
0.166
(0.640)
0.183
0.390
21
Figure 3: The equipartition energy density of radio sources as a function of
redshift. Selected "normal-sized" sources are marked with the full squares.
A number of already known GRGs are marked with crosses, while the newly
detected GRGs (given in Table 5) are marked with the open circles. The
dashed line follows the proportionality p ∝ (1 + z)5. An uncertainty of the
redshift estimate for the sample GRG J1420−0545 is marked with the full
line, cf. the text. The solid curve indicates a relation ueq ∝ (P/V )4/7, cf. the
text
22
|
astro-ph/0503541 | 1 | 0503 | 2005-03-24T14:56:31 | Grid services for the MAGIC experiment | [
"astro-ph"
] | Exploring signals from the outer space has become an observational science under fast expansion. On the basis of its advanced technology the MAGIC telescope is the natural building block for the first large scale ground based high energy gamma-ray observatory. The low energy threshold for gamma-rays together with different background sources leads to a considerable amount of data. The analysis will be done in different institutes spread over Europe. Therefore MAGIC offers the opportunity to use the Grid technology to setup a distributed computational and data intensive analysis system with the nowadays available technology. Benefits of Grid computing for the MAGIC telescope are presented. | astro-ph | astro-ph |
GRID SERVICES FOR THE MAGIC EXPERIMENT
A. FORTI 1, S.R. BAVIKADI 1, C. BIGONGIARI 2,
G. CABRAS 1, A. DE ANGELIS 1, B. DE LOTTO 1,
M. FRAILIS 1, M. HARDT 3, H. KORNMAYER 3,
M. KUNZE 3, M. PIRACCINI 1 AND THE MAGIC
COLLABORATION
1 INFN and Dipartimento di Fisica, Universit´a degli Studi di
Udine, Italy
2 INFN and Dipartimento di Fisica Galileo Galilei, Padova, Italy
3 Institut
fur wissenschaftliches Rechnen, Forschungszentrum
Karlsruhe, Germany
Abstract
Exploring signals from the outer space has become an observational science
under fast expansion. On the basis of its advanced technology the MAGIC
telescope is the natural building block for the first large scale ground based
high energy γ-ray observatory. The low energy threshold for γ-rays together
with different background sources leads to a considerable amount of data.
The analysis will be done in different institutes spread over Europe. There-
fore MAGIC offers the opportunity to use the Grid technology to setup
a distributed computational and data intensive analysis system with the
nowadays available technology. Benefits of Grid computing for the MAGIC
telescope are presented.
1 MAGIC
The MAGIC telescope has been designed to search the sky to discover or observe
high energy γ-rays sources and address a large number of physics questions [1].
Located at the Instituto Astrophysico de Canarias on the island La Palma, Spain,
1
Proceedings of the 6th International Symposium
"Frontiers of Fundamental and Computational Physics", Udine (Italy), Sep. 26-29, 2004
c(cid:13) 2005 Kluwer Academic Publishers, Boston. Manufactured in The Nederlands.
2
Grid services for the MAGIC telescope
at 28◦ N and 18◦ W, at altitude 2300m asl, it is the largest γ-ray telescope in the
world. MAGIC is operating since October 2003, data are taken regularly since
February 2004 and signals from Crab and Markarian 421 was seen. The main
characteristics of the telescope are summarized below:
• A 17m diameter (f/d=1) tessellated mirror mounted on an extremely light
carbon-fiber frame (< 10 tons), with active mirror control. The reflecting
2; reflectivity is larger than 85% (300 - 650nm).
surface of mirrors is 240 m
• Elaborate computer-driven control mechanism.
• Fast slewing capability (the telescope moves 180◦ in both axes in 22s).
• A high-efficiency, high-resolution camera composed by an array of 577 fast
photomultipliers (PMTs), with a 3.9◦ field of view.
• Digitalization of the analogue signals performed by 300 MHz FlashADCs
and a high data acquisition rate of up to 1 KHz.
• MAGIC is the lowest threshold (≈ 30 GeV) IACT operating in the world.
Gamma-rays observation in the energy range from a few tenth of GeV upwards, in
overlap with satellite observations and with substantial improvement in sensitivity,
energy and angular resolution, leads to search behind the physics that has been
predicted and new avenues will open. AGNs, GRBs, SNRs, Pulsars, diffuse photon
background, unidentified EGRET sources, particle physics, darkmatter, quantum
gravity and cosmological γ-ray horizon are some of the physics goals that can be
addressed with the MAGIC telescope.
2 Grid
The idea of computational and data grids dates back to the first half of the 90's.
The vision behind them is often explained using the electric power grid metaphor.
The electric power grid delivers electric power in a pervasive and standardized
way. You can use any device that requires standard voltage and has a standard
plug if you are able to connect it to the electric power grid through a standard
socket. When you use electricity you don't worry were it is produced and how it
is delivered, you just plug your device into the wall socket and use it. Currently
we have millions of computing and storage systems all over the planet connected
through the Internet. What we need is an infrastructure and standard interfaces
capable of providing transparent access to all this computing power and storage
space in a uniform way. This is the concept behind Grid. More precisely, Grid
is a kind of parallel and distributed system that enables the sharing, selection
and aggregation of services of heterogeneous resources distributed across multiple
administrative domains based on their availability, capability, performance, cost,
and users quality-of-service requirements [2]. As network performance has out-
paced computational power and storage capacity, this new paradigm has evolved
to enable the sharing and coordinated use of geographically distributed resources.
A. Forti et al.
3
2.1 VIRTUAL ORGANIZATION
The Virtual organization is an important Grid concept. Grid allows a pool of
heterogeneous resources both within and outside of an organization to be virtual-
ized and form a large, virtual computer. This virtual computer can be used by a
collection of users and/or organizations in collaboration to solve their problems.
The rules governing the participants providing the resources and the consumers
using the resources, as well as the conditions for sharing, dictate the nature of the
virtual organization. Hence, a virtual organization groups people and resources
without worry about their physical location or institute boundaries. For security
reasons, the use of the resources is constrained by an authentication process.
3 Benefits of Grid computing for Magic
The collaborators of the MAGIC telescope are mainly spread over Europe, 18
institutions from 9 countries, with the main contributors (90% of the total) lo-
cated in Germany (Max-Planck-Institute for Physics, Munich and University of
Wuerzburg), Spain (Barcelona and Madrid), Italy (INFN and Universities of Pa-
dova, Udine and Siena).
The geographical distribution of the resources makes the management of the
experiment harder. This is a typical situation for which Grid computing can be of
great help, because it allows researchers to access all the resources in a uniform,
transparent and easy way. The telescope is in operation during moonless nights.
The average amount of raw FADC data recorded is about 500-600 GB/night.
Additional data from the telescope control system or information from a weather
station are also recorded. All these information have to be taken into account in
the data analysis.
The MAGIC community can leverage from Grid facilities in areas like file shar-
ing, Monte Carlo data production and analysis. In a Grid scenario the system can
be accessed through a web browser based interface with single sign-on authentica-
tion method. We can briefly summarize the main benefits given by the adoption
of Grid technology for the MAGIC experiment:
• Presently, users analyzing data must know where to find the required files
and explicitly download them. In a Grid perspective, instead, users don't
care about data location and files replication policies improve access time
and fault tolerance.
• Grid workflow tools can manage the MAGIC Monte Carlo simulation. The
resources from all the members of the MAGIC community can be put to-
gether and exploited by the Grid. Easy access to data production for every
user, or accordingly to the VO policies.
• Analysis tools can be installed on the Grid. They are thus shared and avail-
able for all the users (no need for single installations), moreover, they can
4
Grid services for the MAGIC telescope
Figure 1. MAGIC Monte Carlo Simulation workflow
exploit the facilities of a distributed system.
3.1 MMCS
The MAGIC Monte Carlo simulation workflow is a series of programs which sim-
ulate the properties of different physics processes and detector parts (figure 1):
• CORSIKA: air shower and hadronic background simulation. The output
contains information about the particles and the Cherenkov photons reaching
the ground around the telescope.
• Reflector : simulates the propagation of Cherenkov photons through the at-
mosphere and their reflection in the mirror up to the camera plane. The
input for the Reflector program is the output of CORSIKA.
• StarfieldAdder : simulation of the field of view. It adds light from the non-
diffuse part of the night sky background, or the effect of light from stars, to
images taken by the telescope.
• StarResponse: simulation of NSB (night sky background) response.
• Camera: simulate the behavior of the photomultipliers and of the electronic
It also allows to introduce the NSB (optionally),
of the MAGIC camera.
from the stars and/or the diffuse NSB.
A. Forti et al.
5
3.2 THE PRESENT
The EGEE (Enabling Grids for E-science in Europe) project brings together ex-
perts from 70 organizations and 27 countries with the common aim of building on
recent advances in Grid technology and developing a service grid infrastructure
in Europe which is available to scientists 24 hours-a-day. Recently MAGIC has
became part of the EGEE project. This is the first step for enabling MAGIC on
the Grid. This process migration is a big effort and must be divided in smaller
steps. The first step was chosen to be the migration of the MAGIC Monte Carlo
simulation workflow and it is now working [3] thanks to the effort of the Udine,
Padova and CNAF (Bologna) groups.
4 Conclusions
Grid technologies promise to change the way organizations tackle their complex
computational and data-intensive problems. The vision of large scale resource
sharing is nowadays becoming a reality in many areas. However, it must be realized
that Grid is an evolving field in computer science, where standards and technology
are still being under development to enable this new paradigm. Presently, many
efforts are being made to attract a wide range of new users to the grid. MAGIC has
caught this big opportunity and now another step is made towards the realization
of a wide project that wants to connect the compute and storage resources of
the astroparticle institutions in order to collaborate across the institute borders as
well as across the collaboration borders. This new challenge will be the ASTROPA
Grid project [4].
References
[1] Lorenz, E., New Astron. Rev. vol. 48, 339-344 (2004).
[2] Foster, I., Kesselman, C. (eds.). The Grid: Blueprint for a New Computing Infrastructure.
Morgan Kaufmann, 1999.
[3] M. Piraccini, tesi di Laurea in Fisica Computazionale, Udine 2004.
[4] ASTROPA Grid, proposal by the MAGIC Collaboration to EGEE, 2004.
|
0807.0508 | 1 | 0807 | 2008-07-03T08:55:20 | Measurements of Stellar Properties through Asteroseismology: A Tool for Planet Transit Studies | [
"astro-ph"
] | Oscillations occur in stars of most masses and essentially all stages of evolution. Asteroseismology is the study of the frequencies and other properties of stellar oscillations, from which we can extract fundamental parameters such as density, mass, radius, age and rotation period. We present an overview of asteroseismic analysis methods, focusing on how this technique may be used as a tool to measure stellar properties relevant to planet transit studies. We also discuss details of the Kepler Asteroseismic Investigation -- the use of asteroseismology on the Kepler mission in order to measure basic stellar parameters. We estimate that applying asteroseismology to stars observed by Kepler will allow the determination of stellar mean densities to an accuracy of 1%, radii to 2-3%, masses to 5%, and ages to 5-10% of the main-sequence lifetime. For rotating stars, the angle of inclination can also be determined. | astro-ph | astro-ph | Transiting Planets
Proceedings IAU Symposium No. 253, 2008
c(cid:13) 2008 International Astronomical Union
DOI: 00.0000/X000000000000000X
Measurements of Stellar Properties through
Asteroseismology:
A Tool for Planet Transit Studies
Hans Kjeldsen1, Timothy R. Bedding2 and
Jørgen Christensen-Dalsgaard1
1Danish AsteroSeismology Centre, Department of Physics and Astronomy, University of
Aarhus, Ny Munkegade, DK-8000 Aarhus C, Denmark, email: [email protected] and
2Institute of Astronomy, School of Physics A28, University of Sydney, NSW 2006, Australia.
email: [email protected]
[email protected]
Abstract. Oscillations occur in stars of most masses and essentially all stages of evolution.
Asteroseismology is the study of the frequencies and other properties of stellar oscillations, from
which we can extract fundamental parameters such as density, mass, radius, age and rotation
period. We present an overview of asteroseismic analysis methods, focusing on how this tech-
nique may be used as a tool to measure stellar properties relevant to planet transit studies. We
also discuss details of the Kepler Asteroseismic Investigation -- the use of asteroseismology on
the Kepler mission in order to measure basic stellar parameters. We estimate that applying as-
teroseismology to stars observed by Kepler will allow the determination of stellar mean densities
to an accuracy of 1%, radii to 2 -- 3%, masses to 5%, and ages to 5 -- 10% of the main-sequence
lifetime. For rotating stars, the angle of inclination can also be determined.
Keywords. stars: interiors, stars: oscillations, stars: rotation, methods: data analysis
1. Introduction
Asteroseismology -- the study of stellar oscillations -- is a relatively new and growing
research field in astrophysics. The analysis of frequencies and other properties of stellar
oscillations allows us to constrain fundamental parameters of stars such as density, mass,
radius, age, rotation period and chemical composition.
Oscillations are found in stars of most masses and essentially all stages of evolution.
The amplitudes and phases are controlled by the energetics and dynamics of the near-
surface layers. The frequencies are determined by the internal sound-speed and density
structure of the star, as well as rotation and (in some cases) magnetic fields. Observation-
ally, the frequencies can be determined with exceedingly high accuracy compared to any
other quantity relevant to the internal properties of the stars. Analysis of the observed
frequencies, including comparison with computed stellar models, allows determination of
the properties of the stellar interiors and tests of the physics used in the model computa-
tion, applied under extreme conditions that cannot be matched in terrestrial laboratories.
Rotation induces fine structure in the frequency spectrum, in the form of rotational
splitting. The observed frequencies are determined by averages over the stellar interior
of the rotation rate, which in general varies with position within the star. By comparing
with independent determinations of the surface rotation rate, or from rotational splittings
for a sufficient broad variety of modes, information about this variation can be obtained.
Some of the basic parameters that may be obtained from asteroseismology are crucial
for understanding the fundamental properties of exoplanet systems. Hence, asteroseis-
1
8
0
0
2
l
u
J
3
]
h
p
-
o
r
t
s
a
[
1
v
8
0
5
0
.
7
0
8
0
:
v
i
X
r
a
2
Kjeldsen, Bedding & Christensen-Dalsgaard
mology will potentially be an excellent tool for characterizing planet transit systems and,
specifically, it will be able to measure stellar radii to a relative accuracy of 2 -- 3%.
For a recent review on the present state of asteroseismology, see Aerts et al. (2008)
and Bedding & Kjeldsen (2008). In Figure 1 we show examples of power spectra for four
different stars. In each case, we can clearly see the excess energy arising from stellar
oscillations. The detailed properties of those oscillations form the observational basis for
the asteroseismic analysis.
r
e
w
o
P
0.1
1.0
Frequency (mHz)
Figure 1. Power spectra of time series observed in radial velocity for the four stars ξ Hydrae
(G7 III; Frandsen et al. 2002), Procyon (F5 IV; Arentoft et al. 2008), α Centauri A (G2 V; Bed-
ding et al. 2004) and α Centauri B (K2 V; Kjeldsen et al. 2005). The oscillation periods for
the four stars are: around 3 -- 4 hours for ξ Hydrae, 20 -- 25 minutes for Procyon, 7 minutes for
α Cen A and 4 minutes for α Cen B. These differences, and those in the detailed structure of
the power spectra, reflect the differences in stellar properties (radius, mass, surface temperature
and age). Note that the vertical scales are normalized -- the actual amplitudes decrease from top
to bottom in the figure.
Asteroseismology, a tool for transit studies
3
2. The relation between the stellar properties and the frequencies
To obtain information about stellar properties and the underlying physics, the observed
oscillation frequencies must be compared with those computed from models. Major un-
certainties affect the treatment of hydrodynamical processes in stellar interiors. These
include convective motions, with possible overshoot into the surrounding convectively
stable regions, and flows and instabilities related to internal rotation. Furthermore, the
evolution of rotation as a star evolves, with internal redistribution and possibly surface
loss of angular momentum, has potentially substantial -- and highly uncertain -- effects
on the evolution of the star. Despite those uncertainties in the knowledge of some of the
detailed physics, we are able to obtain quite detailed basic properties for a number of
general stellar parameters such as density and the amount of hydrogen in the core. The
frequencies of stellar p-mode oscillations are related to the sound travel time across the
star. Since the sound speed, c, is given by
c2 =
Γ1 P
ρ
≃
5
3
kBT
µmu
,
(2.1)
(where the approximation assumes an ideal gas), we are basically measuring the average
ratio between pressure and density in the stellar interior.
The exact frequencies of stellar oscillations depend on the detailed structure of the star
and on the physical properties of the gas. However, one may use asymptotic theory to
0(cid:542)(cid:39)
13(cid:303)(cid:542)
02(cid:303)(cid:542)
Figure 2. The power spectrum of oscillations in the Sun. The data are full-disk radial velocity
measurements obtained over 30 days using the GOLF instrument on board the SoHO spacecraft
(Ulrich et al. 2000, Garc´ıa et al. 2005). The inset shows the details of the p-mode structure, and
we indicate the so-called large and small separations, which contain information on the basic
stellar properties.
4
Kjeldsen, Bedding & Christensen-Dalsgaard
derive a simple relation for the mode frequencies (see, for example, Christensen-Dalsgaard
2004). Each oscillation mode is characterized by three integers: the radial order n, the
angular degree l and the azimuthal order m. The results of the asymptotic analysis for
a non-rotating star give the mode frequencies as
νn,l = ∆ν(n + 1
2 l + ǫ) − l(l + 1)D0.
(2.2)
Here, ∆ν (the so-called large separation) depends on the average stellar density, D0 is
sensitive to the sound speed near the core and ǫ is sensitive to the surface layers. It
is conventional to define δν02, the so-called small separation, as the frequency spacing
between adjacent modes with l = 0 and l = 2. These separations are indicated for the
Sun in Figure 2, together with the similar quantity δν13. We can further define δν01
to be the amount by which l = 1 modes are offset from the midpoint between the
l = 0 modes on either side. If the asymptotic relation holds exactly, then it follows that
D0 = 1
2 δν01 = 1
6 δν02 = 1
10 δν13.
In practice, the asymptotic relation does not hold exactly, even for the Sun. For ex-
ample, the large separation depends on l. However, we can define average values for ∆ν
and δν02, and then can calculate how those parameters depend on the stellar properties
R(cid:71)
R
%3(cid:100)
M(cid:71)
M
%5(cid:100)
(cid:302) Cen B
The Sun
(cid:302) Cen A
(
Age(cid:71)
Age
MS
)
(cid:100)
%10
Figure 3. The so-called asteroseismic HR-diagram (Christensen-Dalsgaard 1993), where the
axes are the large and small frequency separations (∆ν and δν02). The solid lines are evolution-
ary tracks (for fixed mass) and the dashed lines show constant core-hydrogen content, which
is related to the age. Based on simulations we have estimated the uncertainties for the mea-
surements of mass, radius and stellar age using the seismic HR-diagram to be 5% for the mass,
2 -- 3% for the radius and 5 -- 10% for the age (relative to the total time on the main sequence).
The mean densities can be measured to 1%. The positions of the Sun and α Centauri A and B
are indicated, and it can be seen that α Cen A is more evolved than the Sun while α Cen B is
less evolved, in agreement with traditional modeling.
Asteroseismology, a tool for transit studies
5
and the stellar evolutionary stage. An example of such a calculation is shown by the lines
in Figure 3. In this figure we also show the measured position of the Sun, together with
that of α Centauri A (Bedding et al. 2004) and α Centauri B (Kjeldsen et al. 2005). This
diagram confirms that α Cen A has a higher mass than the Sun and is more evolved,
while α Cen B has a lower mass than the Sun and is less evolved.
The exact positions of the tracks shown in Figure 3 will depend on the detailed physical
properties, including chemical composition. However, the basic stellar parameters can still
be estimated with some robustness. We have used simulations to estimate the accuracy
with which various parameters can be measured, assuming the stellar temperature is
already known to within 150 -- 200 K and the heavy-element abundance is known to within
a factor of two. We find that
• stellar mean densities can be measured to 1%,
• stellar radii can be measured to 2 -- 3%,
• stellar masses can be measured to 5% and
• stellar ages can be measured to 5 -- 10% of the main-sequence lifetime.
l=2, n=12
m=-2,0,2
l=0, n=13
r
e
w
o
P
e
v
i
t
a
e
R
l
1.2
1.0
0.8
0.6
0.4
0.2
0.0
1.940
1.945
1.950
Frequency (mHz)
1.955
1.960
Figure 4. A close-up of the power spectrum of solar oscillations, from full-disk radial velocity
measurements. The data are based on 805 days of observing the Sun using the GOLF instrument
on board the SoHO spacecraft (Ulrich et al. 2000, Garc´ıa et al. 2005). We show a small region
of the power spectrum just below 2 mHz. In this region one can see the power from modes with
l = 2, n = 12 and l = 0, n = 13. The l = 2 multiplet has 5 components (m = 0, ±1, ±2) but, due
to the inclination of the rotation axis of the Sun, only modes with m = −2, 0 and 2 are visible.
The strength of the individual m-components can be used to measure the inclination axis of the
stellar rotation. Modes with l = 0 have only m = 0 and are therefore not split by rotation. The
upper curve is a smoothed version of the power spectrum, shifted vertically for clarity.
6
Kjeldsen, Bedding & Christensen-Dalsgaard
3. Stellar Rotation
Equation 2.2 applies to a non-rotating star, for which the frequency of each mode
depends only on the radial order n and the angular degree l. For a rotating star, the
frequencies also depend on the azimuthal degree, m (which takes on values of m =
0, ±1, . . . , ±l). Provided the rotation is slow, the frequencies are well approximated by
νn,l = ∆ν(n + 1
2 l + ǫ) − l(l + 1)D0 + m∆νROT.
(3.1)
Here, ∆νROT is related to the inverse of average of the internal rotation period and we
should therefore be able to infer stellar rotation periods directly from the oscillation
frequencies.
In principle, the inclination of the rotation axis can also be determined because it
affects the relative amplitudes of the different azimuthal degrees. This effect is discussed in
detail by Gizon & Solanki (2003). For example, they show that for intensity observations
(oscillations detected in photometry) the dipole multiplets (l = 1, m = −1, 0, 1) will
have relative mode powers given by
Pl=1,m=0 = cos2 i ,
Pl=1,m=±1 = 1
2 sin2 i .
(3.2)
(3.3)
In Figures 4, 5 and 6 we show examples of rotational splitting for the Sun. The splitting
of the l = 1 and 2 modes is clearly seen. The relative strengths of the different m
components is a result of the observed inclination of the solar rotation axis. Since we
observe the Sun in the equatorial plane, we have almost no sensitivity to the modes with
l=1, m=0 and those with l=2, m=±1. The structure of the power spectrum therefore
l=2, n=20
m=-2,0,2
l=0, n=21
r
e
w
o
P
e
v
i
t
a
e
R
l
1.2
1.0
0.8
0.6
0.4
0.2
0.0
3.020
3.025
Frequency (mHz)
3.030
3.035
3.040
Figure 5. Same as Figure 4, but around a frequency of 3 mHz. We see the multiplet with
n = 20 and l = 2, and the singlet with n = 21 and l = 0.
Asteroseismology, a tool for transit studies
7
indicates that the solar interior is rotating in the surface equatorial plane of the Sun,
approximately corresponding to the plane of the ecliptic.
4. The Kepler Asteroseismic Investigation
The NASA Kepler mission (Borucki et al. 2008) will fly a wide-field Schmidt camera
with 0.95m aperture, staring at a single field continuously for at least 3.5 years. Although
the mission's principal aim is to locate transiting extrasolar planets, it will provide an
unprecedented opportunity to make asteroseismic observations of a wide variety of stars.
This will give the opportunity to measure global properties and internal structure for
a large number of stars across a broad range of different types. Plans are now being
developed to exploit this opportunity to the fullest. In particular, asteroseismic analysis
of Kepler data will provide an accurate determination of the radius for a large fraction
of the stars found to host planetary systems, as determined from the transit analysis of
the Kepler data, as well as estimates of the ages of the systems. Through investigation of
a broad range of stars, the asteroseismic investigation will also substantially improve our
understanding of general stellar evolution, and hence strengthen the use of such modelling
to further constrain the properties and evolution of the stars and systems investigated
in the Kepler extra-solar planet programme. The purpose of the Kepler Asteroseismic
Investigation (e.g., Christensen-Dalsgaard et al. 2007) is to ensure that full use is made
of this potential to benefit the Kepler investigations of extra-solar planetary systems.
Kepler will give the possibility of observing at two cadences (1 minute and 30 minutes).
Due to the high frequency of the stellar oscillations Kepler will not be able to make
l=1, n=15
m=-1,1
r
e
w
o
P
e
v
i
t
a
e
R
l
1.5
1.0
0.5
0.0
2.288
2.290
Frequency (mHz)
2.292
2.294
Figure 6. Same as Figure 4, but at a frequency just below 2.3 mHz. In this region one can see
the power from the triplet with l = 1, n = 15, m = 0, ±1. Due to the inclination of the rotation
axis of the Sun, only modes with m = −1 and 1 are visible.
8
Kjeldsen, Bedding & Christensen-Dalsgaard
asteroseismic observations of all the 170,000 stars observed in long cadence mode (30
minute cadence). We will therefore need a target selection programme in order to optimize
the asteroseismic part of the Kepler programme. Based on simulations, we have estimated
that seismology can be done on unevolved main-sequence stars down to magnitude V =
12.5, and on evolved main-sequence stars (which have higher amplitudes) down to V =
13.5. For subgiant stars we may reach even fainter. It should be noted that Kepler will
be the first mission where asteroseismology will be applied to a large number of planet-
hosting stars.
Finally, we note that accurate measurements of variations in the light travel times for
planets orbiting classical pulsating stars (the pulsations are here used as accurate clocks)
may provide a unique tool for the Kepler mission to detect a number of additional plan-
ets. The technique requires accurate measurements of phase variations of the pulsations
(better than seconds) and the method will be most sensitive to long-period planets (1 -- 2
years). For details on this technique we refer to Silvotti et al. (2007).
References
Aerts, C., Christensen-Dalsgaard, J., Cunha, M. & Kurtz, D. W., 2008, Solar Phys., in the press
[arXiv:0803.3527]
Arentoft, T., Kjeldsen, H., Bedding, T. R., et al. 2008, ApJ, in the press
Bedding, T. R. & Kjeldsen, H. 2008, in: G. van Belle (ed.), 14th Cambridge Workshop on Cool
Stars, Stellar Systems, and the Sun, ASP Conf. Ser., Vol. 384 (San Francisco: ASP), 21
Bedding, T. R., Kjeldsen, H., Butler, R. P., et al. 2004, ApJ 614, 380
Borucki, W. et al. 2008, these Proceedings
Rotation:
Period and inclination
Mass
Age
Radius
Stellar activity
Figure 7. Using asteroseismology we are able to measure detailed basic properties of stars that
have a transiting planet. Asteroseismology is expected to become a crucial tool for the Kepler
mission when measuring stellar parameters.
Asteroseismology, a tool for transit studies
9
Christensen-Dalsgaard, J. 1993, in: T. M. Brown (ed.), Proc. GONG 1992: Seismic investigation
of the Sun and stars, ASP Conf. Ser., Vol. 42 (San Francisco: ASP), 347
Christensen-Dalsgaard, J. 2004, Solar Phys. 220, 137
Christensen-Dalsgaard, J., Arentoft, T., Brown, T. M., Gilliland, R. L., Kjeldsen, H., Borucki,
W. J. & Koch, D. 2007, Comm. in Asteroseismology 150, 350
Frandsen, S., Carrier, F., Aerts, C., et al. 2008, A&A 394, L5
Garc´ıa, R. A., Turck-Chi`eze, S., Boumier, P., et al. 2005, A&A 442, 385
Gizon, L. & Solanki, S. K. 2003, ApJ 589, 1009
Kjeldsen, H., Bedding, T. R., Butler, R. P., et al. 2005, ApJ 635, 1281
Silvotti, R., Schuh, S., Janulis, R., et al. Nature 449, 189
Ulrich, R. K., Garc´ıa, R. A., Robillot, J.-M. et al. 2000, A&A 364, 799
|
astro-ph/0210159 | 3 | 0210 | 2003-02-19T20:07:48 | Velocity statistics from spectral line data: effects of density-velocity correlations, magnetic field, and shear | [
"astro-ph"
] | In a previous work Lazarian and Pogosyan suggested a technique to extract velocity and density statistics, of interstellar turbulence, by means of analysing statistics of spectral line data cubes. In this paper we test that technique, by studying the effect of correlation between velocity and density fields, providing a systematic analysis of the uncertainties arising from the numerics, and exploring the effect of a linear shear. We make use of both compressible MHD simulations and synthetic data to emulate spectroscopic observations and test the technique. With the same synthetic spectroscopic data, we also studied anisotropies of the two point statistics and related those anisotropies with the magnetic field direction. This presents a new technique for magnetic field studies. The results show that the velocity and density spectral indices measured are consistent with the analytical predictions. We identified the dominant source of error with the limited number of data points along a given line of sight. We decrease this type of noise by increasing the number of points and by introducing Gaussian smoothing. We argue that in real observations the number of emitting elements is essentially infinite and that source of noise vanishes. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 29 October 2018
(MN LATEX style file v2.2)
Velocity statistics from spectral line data: effects of density-velocity
correlations, magnetic field, and shear
A. Esquivel,1⋆, A. Lazarian,1⋆ D. Pogosyan,2⋆ J. Cho1⋆
1Astronomy Department, University of Wisconsin-Madison, 475 N. Charter St., Madison, WI 53706, USA
2Department of Physics,University of Alberta, Edmonton, Alberta T6G 2J1, Canada
Draft Version 29 October 2018
ABSTRACT
In a previous work Lazarian and Pogosyan suggested a technique to extract velocity and den-
sity statistics, of interstellar turbulence, by means of analysing statistics of spectral line data
cubes. In this paper we test that technique, by studying the effect of correlation between ve-
locity and density fields, providing a systematic analysis of the uncertainties arising from the
numerics, and exploring the effect of a linear shear. We make use of both compressible MHD
simulations and synthetic data to emulate spectroscopic observations and test the technique.
With the same synthetic spectroscopic data, we also studied anisotropies of the two point
statistics and related those anisotropies with the magnetic field direction. This presents a new
technique for magnetic field studies. The results show that the velocity and density spectral
indices measured are consistent with the analytical predictions. We identified the dominant
source of error with the limited number of data points along a given line of sight. We decrease
this type of noise by increasing the number of points and by introducing Gaussian smoothing.
We argue that in real observations the number of emitting elements is essentially infinite and
that source of noise vanishes.
Key words: turbulence -- ISM: general, structure -- MHD -- radio lines: MHD.
1 INTRODUCTION
The very large Reynolds numbers (defined as the ratio of the in-
ertial force to the viscous force acting on a parcel of gas) clearly
suggest that the interstellar medium is turbulent. Understanding of
this turbulence is crucial to the correct and complete description
of many physical processes that take place in the ISM: molecular
cloud dynamics, star formation, heat transfer, magnetic reconnec-
tion, accretion disks, cosmic ray propagation and diffusion, just to
mention a few.
To adequately describe turbulence we need to use statistical
methods, to extract the underlying regularities and reject incidental
details. So far, studies of turbulence statistics have been most suc-
cessful using interstellar scintillations, and have provided impor-
tant information of the density statistics on scales 108 − 105cm,
(see Narayan & Goodman 1989; Spangler & Gwinn 1990). These
studies are done with the ionised media, and are restricted to the
density fluctuations. We expect to observe power-law power spec-
trum of density in a turbulent medium, however density fluctu-
ations alone fail to distinguish between fossil and active turbu-
lence. The distribution of sizes of sand grains in a beach follows
a power-law but it is not turbulence. Therefore would be better to
study a more direct and dynamically important statistic: velocity
⋆ E-mail:
[email protected]; [email protected]
[email protected];
[email protected];
fluctuations. Studies of the neutral media have included velocity
information, e.g. using spectral line widths, centroids of veloc-
ity (Miesch & Bally 1994; Scalo 1987); or more fancy statistics
(Heyer & Schloerb 1997; Rosolowsky et al. 1999; Brunt & Heyer
2002a,b).
Turbulence statistics is a very important link between the-
ory and observations. Recent breakthroughs in turbulence studies
have been possible with the use of scaling laws (see reviews by
Cho & Lazarian 2003; Cho, Lazarian & Vishniac 2003, and refer-
ences therein). Scaling laws predictions can be confronted with ob-
servations.
A discussion of various approaches to the study of turbulence
with spectral line data can be found in Lazarian (1999). In particular
the problem of the contribution of velocity and density fluctuations
to the emissivity statistics was addressed in Lazarian & Pogosyan
(2000, hereafter LP00), where analytical description of the power
spectrum of the emissivity in velocity channels was obtained. This
study provides a technique to extract statistics of density and ve-
locity from spectral line data cubes. H I has been chosen as a test
case because it was possible to disregard self-absorption, and con-
sider the emissivity linearly depending on the density (see review
in Lazarian, Pogosyan & Esquivel 2002). The technique, termed as
velocity-channel analysis (or simply, VCA), is based on the vari-
ations of the power spectra in velocity channels with the change
of velocity resolution. Predictions in LP00 have been confirmed
through observations (Stanimirovi´c & Lazarian 2001). The first
2
Esquivel et al.
tests using numerical data in Lazarian et al. (2001) also confirmed
the predictions in LP00. However they did not address the problems
of numerical noise, effects of magnetic field and galactic shear.
Moreover the study of velocity-density correlations was performed
over a very limited dynamical range.
In this work we subject the predictions in LP00 to further
scrutiny. In particular, we systematically analyse the effects of
velocity-density correlations, and unlike Lazarian et al. (2001) we
include a study of the effect of shear. In addition we address the is-
sue of an anisotropic turbulent cascade, with their implications on
the VCA. We also analyse the sources of numerical noise.
In §2 we review the basic problem, we then (§3) test the
extraction of the spectral indices using the VCA, and obtain the
density-velocity correlations. In §4 we study the effect of shear in
the VCA. The anisotropic cascade, and a possible new method to
extract information of the magnetic field direction can be found in
§5. And we finally provide a summary of our results (§6).
2 STATISTICS OF SPECTRAL LINE DATA AND VCA
Spectral line observations contain important information of den-
sity and velocity. The problem is that both velocity fluctuations and
density fluctuations contribute to the intensity of a spectral line at a
given velocity, and their separation is far from trivial.
2.1 3D turbulence statistics
Turbulence, due to its stochastic nature is studied using statis-
tical tools, such as 3D correlation and structure functions (see
Monin & Yaglom 1975). For instance, the density correlation func-
tion can be expressed as
ξ(r) = hρ(x)ρ(x + r)i,
where x is the spatial position (xyz in Cartesian coordinates), r
is a spatial separation (or 'lag'), and the angular brackets denote
an average to be performed over all x space. The power spectrum
provides an alternative description and is related to the correlation
function (CF) as
(1)
P (k) = Z dr ei k ·r ξ(r),
(2)
where k is the wave-number, related to the scale under study like
k = 2π/r, and the integration is performed over all 3D space. For
power-law statistics, the N dimensional power spectrum (PN ∝
kn) and the correlation (ξN ∝ rm) have indices related as n =
−N − m, in other words1
(spectral index) = (−spatial dimensions − CF index). (3)
To describe velocity fields is customary to use structure functions:
SF (r) = h[v(x) − v(x + r)]2i.
They provide a more formal
treatment of velocity fields
(Monin & Yaglom 1975). In the case of isotropic velocity fields
it differs from the correlation function only by a constant. There-
fore for power-law statistics it has the same spectral index m as
(4)
1 Notice that the spectral indices contain the sign. This notation is unfor-
tunate (contrary to a conventional spectral index of the form PN ∝ k−n,
with n positive). However, it was introduced in LP00, and for consistency
we adhere to it.
the correlation function (i. e. SF ∝ rm). For instance, Kolmgorov
turbulence scales as v ∝ r1/3, which corresponds to m = 2/3.
2.2 Statistics in position-position-velocity (PPV) space
In spectroscopic observations we don't observe the gas distribution
in real space coordinates x ≡ (x, y, z). Instead the intensity of the
emission in a given spectral line is measured towards some direc-
tion in the sky, given by X ≡ (x, y), and at a given line-of-sight
(LOS) velocity v. Thus, the coordinates of observational data are
(X, v). Therefore, observational data are contained in, so called,
position-position-velocity (or simply 'PPV') cubes. If we identify
the z coordinate with the LOS, then the relation between the real
space and PPV descriptions is that of a map (X, z) → (X , v).
At a given position x the LOS component of the velocity v can
be decomposed as a regular flow velocity ureg(x), plus a thermal
component vthermal and a turbulent component u(x). This way the
observed Doppler shifted atoms follow a Maxwellian distribution
of the form
φv(x)dv =
1
(2πβ)1/2 exp(cid:26)−
[v − vreg(x) − u(x)]2
2β
(cid:27) dv,
(5)
where β = κBT /m, m is the atomic mass, T the gas temperature,
and κB the Maxwell-Boltzmann constant.
The density of emitters in PPV space ρs(X, v) can be ob-
tained by integration along the LOS:
ρs(X, v)dXdv = (cid:20)Z dz ρ(x)φ(x)(cid:21) dXdv,
(6)
where ρ(x) is the mass density of the gas in spatial coordinates.
This expression just counts the number of atoms along the LOS, at
a given position X with a z component of velocity in the interval
[v, v + dv], and the limits of integration are defined by the extent
of spatial distribution of emitting gas.
The statistics available in PPV are the correlation functions,
ξs ≡ hρs(X1, v1)ρs(X2, v2)i, and the power spectra (this can be
1D, 2D or 3D) of emissivity fluctuations. Relations between this
statistics and the underlying velocity and density spectra were es-
tablished in LP00, where the 2D power spectrum in velocity chan-
nels was used. This spectrum can be directly obtained with radio-
interferometric observations (see Lazarian 1995).
2.3 Velocity-channel analysis (VCA)
This technique was developed in LP00 to extract the velocity and
density spectral indices of turbulence from spectral line data. LP00
provide analytical predictions for the emissivity power spectral in-
dices in velocity channels, as a function of the velocity resolution
employed. The relative contribution to the intensity in a velocity
channel (also referred as 'slice') to the total intensity fluctuations
changes in a regular fashion with the width of the velocity slice.
This is easy to understand qualitatively since more velocity fluctua-
tions within a channel are averaged out as we increase the thickness
of the channel, decreasing the relative contribution of velocity.
The notation that we will use is such that: n is the 3D density
spectral index (Pρ ∝ kn); µ the 3D velocity spectral index (Pv ∝
kµ); γ corresponds to the spectral index in velocity slices (Pn ∝
kγ); and m is the structure function index of the velocity. From eq.
(3) is clear that
µ = −3 − m.
(7)
Velocity statistics from spectral line data
3
Table 1. A summary of analytical results derived in LP00.
Slice
thickness
Shallow 3-D density
Pn ∝ kn, n > −3
Steep 3-D density
Pn ∝ kn, n < −3
2-D intensity spectrum for thina slice
2-D intensity spectrum for thickb slice
2-D intensity spectrum for very thickc slice
∝ kn+m/2
∝ kn
∝ kn
∝ k−3+m/2
∝ k−3−m/2
∝ kn
a channel width < velocity dispersion at the scale under study
b channel width > velocity dispersion at the scale under study
c substantial part of the velocity profile is integrated over
For a given 3D density power spectrum two distinct regimes
are present when: a) n > −3 ('shallow' density case), and (b)
n < −3 ('steep' density case). A summary of the predictions in
LP00 for the intensity spectral index in 2D velocity channels is
given in Table 1. In both cases (shallow or steep) the power-law
index 'gradually steepens' with the increase of the velocity slice
thickness. In the thickest velocity channels all the velocity infor-
mation is averaged out and we naturally get the density spectral
index n. On the other hand, the velocity fluctuations dominate in
thin channels. For instance, in the case of a steep density (n < −3)
the structure seen in the velocity channels is determined only by ve-
locity fluctuations (is not explicitly dependant on the index n). This
immediately sends a warning against identifying structures in PPV
as actual density enhancements ('clouds'). The spectral velocity in-
dex can be obtained if we combine the density velocity index from
the very thick channels and the structure function index m from
thin channels (using Table 1 and eq. 7). Notice that the notion of
thin and thick slices depends on the turbulence scale under study,
and the same slice can be thick for small-scale turbulent fluctua-
tions and thin for large scale ones. The formal criterion for the slice
to be thick is that the dispersion of turbulent velocities on the scale
studied should be less than the velocity slice thickness. Otherwise
the slice is thin.
From Table 1 one may observe that for any given density spec-
tral index, the spectrum of fluctuations becomes shallower as the
velocity field gets steeper. This is somewhat counterintuitive. But
can be understood as follows: a steeper velocity index corresponds
to relative more power on the largest scales. A steeper velocity
power-spectrum means a larger velocity difference across scales,
and correspondingly to a larger velocity dispersion. In PPV space,
large eddies, because their velocity dispersion, will be spread over
a larger velocity range than smaller eddies. In individual channels,
the number of emitters associated with large eddies will decrease
relative to that of small eddies as the velocity contrast increases.
Therefore, in velocity slices, the power at small k numbers (large
scales, from large eddies) decreases for a steeper velocity, render-
ing in a shallower power-spectrum.
In following sections we perform tests of the VCA using nu-
merical data, and take a step further exploring the sources of uncer-
tainty of and the effects of shear in the VCA.
3 TESTING THE VCA
Numerical simulations provide good means of theory testing. Using
numerical data, we can measure directly the velocity and density
statistics, and we can produce synthetic spectra to analyse them
as an observer would do. However, we need to be aware of the
differences of synthetic and real data. For instance, only a limited
number of points is available with numerics. This, as we show later,
results in the most important source of uncertainty.
3.1 The data
Along the paper we make use of three types of data to construct the
PPV cubes and analyse the power spectrum:
(i) MHD simulations. We use a 3rd-order hybrid essentially-
non-oscillatory (ENO) code to simulate fully developed turbulence.
It was performed on a 2163 Cartesian grid, it is isothermal, com-
pressible, and with a mean magnetic field in the x direction. To
reduce spurious oscillations near shocks, we combine two ENO
schemes. When variables are sufficiently smooth, we use the 3rd-
order weighted ENO scheme (Jiang & Hu 1999) without character-
istic mode decomposition. When opposite is true, we use the 3rd-
order Convex ENO scheme (Liu & Osher 1998). We use a three-
stage Runge-Kutta method for time integration. We solve the ideal
MHD equations in a periodic box:
∂v
∂t
+ v · ∇v + ρ−1∇(c2
∂ρ
∂t
+ ∇ · (ρv) = 0,
sρ) − (∇ × B) × B/4πρ = f ,
∂B
∂t − ∇ × (v × B) = 0,
(8)
with ∇ · B = 0 and an isothermal equation of state. Here f is a
random large-scale driving force, ρ the density, cs the sound speed,
v the velocity, and B the magnetic field. The rms velocity δV is
maintained to be approximately unity, so that v can be viewed as
the velocity measured in units of the rms velocity of the system,
and B/√4πρ as the Alfv´en speed in the same units. The time t
is roughly in units of the large eddy turnover time (∼ L/δV ) and
the length in units of L, the scale of the energy injection. The mag-
netic field consists of the uniform background field and a fluctuat-
ing field: B = B 0 + b. The Alfv´en velocity of the mean field is
roughly the same as the rms velocity. The average Mach number is
∼ 2.5.
The outcomes are the density and velocity fields. The PPV cubes
constructed with this data have a correlation between the velocity
and the density fields which is self-consistent with MHD evolution.
These and are used to test the applicability of the VCA where the
basic assumption was to disregard such correlation.
(ii) 'Modified' data. Due to numerical limitations, the power
spectrum of the simulations mentioned above doesn't show exact
power-law behaviour over the desired inertial range, making diffi-
cult to test the VCA, which deals with power-law spectra. For that
4
Esquivel et al.
reason we apply the same procedure as in Lazarian et al. (2001) to
modify the power spectrum to follow a power-law preserving most
of the phase correlations. The procedure consists in replacing the
amplitudes of the Fourier transform of the data so they follow a
power-law while keeping the phases intact (were most of the spa-
tial information is). The set of simulations used here show a larger
well-developed inertial range than those in Lazarian et al. (2001);
therefore the 'correction' introduced is smaller, especially for the
−11/3.
(iii) Gaussian fields. We use a standard procedure to gen-
erate Gaussian fields with prescribed power-law spectrum (see
Bond, Kofman & Pogosyan 1996). We generate a Fourier represen-
tation of the field
(9)
F (k) = X k ak P (k)1/2 eikr,
where ak are independent Gaussian variables with dispersion 1,
and a mean of 0 for velocity fields or 1 for density fields; in our
notation the wave-number k = 2πL/λ, where λ is the wavelength
associated with k, and L is the box size. To obtain the field in con-
figuration space a Fourier transform is applied. Velocity and density
fields generated this way are entirely uncorrelated. However the fa-
cility of generating as many cubes as we want in a short amount of
time is useful when we want to improve the statistics of our sample.
Comparing (ii) and (iii) allows to test the importance of
density-velocity correlations.
With the data cubes mentioned above we produced synthetic
spectra (i. e. PPV 'cubes'2) to analyse them using VCA. The proce-
dure to construct the PPV cubes is as follows: We generate a matrix
of the same dimensions in X as the x data and a given number of
velocity bins (velocity channels) as the third dimension. The ve-
locity bins are equally spaced ranging from the minimum value of
the velocity (vmin) to the maximum value of the velocity (vmax)
in the whole velocity data cube. We then sum all the elements of
the density cube in a position X that correspond to each veloc-
ity bin. This way we simulate observational data, neglecting self
absorption, and assuming that the thermal width of the emitting
gas is always smaller than the thickness of the velocity channels.
An introduction of an intrinsic width (equivalent to thermal) to the
emitters is discussed later.
3.2 Density and velocity correlations
LP00 considered density and velocity either uncorrelated or cor-
related to maximal degree. What would happen to intermediate
cases? We computed the correlation of the density and the velocity
fields as
C(r) = hρ(x)u(x + r) · r/ri
(10)
σρσu
.
Where r is the spatial separation ('lag'), x is the spatial position,
ρ the density field, u is the velocity field, σρ and σu are the stan-
dard deviations of the density and velocity fields respectively. The
average is to be performed over all x space and angles of r. For
numerical economy Lazarian et al. (2001) used Fourier transform
techniques to obtain similar information (although restricted to the
LOS velocity component). Here this is done directly using equation
(10), which is applied to the original output from the MHD simu-
lations. We found moderate correlations, the maximum correlation
2 Strictly speaking they are parallelepipeds as the sizes need not be the
same in all directions.
Figure 1. Correlation of the density and velocity fields. (a) the average pro-
jection of the cross-correlation on the x axis (parallel to the mean magnetic
field), (b) the average projection of the correlation on the y axis (perpendic-
ular to the mean magnetic field), (c) the average projection of the correla-
tion on the z axis (perpendicular to the mean magnetic field, and (d) average
correlation as function of the spatial separation r.
(in absolute value) was of ∼ 0.22. This correlation is shown in Fig.
1, where we plotted the average projection of the cross-correlation
on the x, y, z axis and the average correlation as a function of the
separation r. Clearly the velocity and density correlations are very
different along and perpendicular to the magnetic field direction.
Some implications of this anisotropy are discussed in §5.
3.3 Applying the VCA
In this subsection we will start the test of the VCA generating sev-
eral data cubes with known ('modified') spectral indices, and ar-
bitrary velocity resolutions that will be shown to range from very
thick, to excessively thin. This is to check the general trends pre-
dicted in LP00. Then with a more careful choice of the velocity
thickness of the channel, we will narrow down the measured spec-
tral indices to check the numerical value with the analytical expec-
tation. A more detailed analysis of the sources of error introduced
with numerics, with an example case of Kolmogorov indices, fol-
lows in the next subsection.
We produced density data cubes with spectral indices modi-
fied to −2.5 and −11/3, LOS velocity cubes with indices of −3.3,
−11/3 (Kolmogorov), and −4.0. With each velocity and the den-
sity fields we constructed PPV cubes with velocity resolutions,
ranging from 150 to 1 number of channels. The 2D power spectra
in velocity channels (averaged over all velocity channels for each
PPV cube) is presented in Fig. 2.
We should stress, that whether a slice in velocity space can
be considered thin or thick depends not only on the slice width
δv = (vmax − vmin)/N = ∆v/N (where N is the number
of channels) but on the scale as well. For power-law statistics the
squared dispersion, σ2
L, over the box size, L, equals CLm (here m
is the velocity structure function index related to the spectral index
as described in eq 3); similarly the velocity dispersion squared, σ2
r ,
over the scale r is Crm. Therefore the criteria for a channel to be
considered thin (δv < σr) from the largest scale L up to the scale
r translates into N > ∆v
power-law result for a slice of a given width. Indeed, the largest
scales, on the order of the size of the cube, are almost always in a
r(cid:1)m/2. Thus, we do not expect a pure
σL (cid:0) L
Velocity statistics from spectral line data
5
Figure 2. Power spectra in velocity channels for simulated observations with density spectral indices of n = −2.5 and n = −11/3, with velocity spectral
indices of µ = −3.3, 11/3, and −4 (as labelled in the title of each plot). In each panel the dotted lines correspond to the power-spectra in velocity channels
for PPV cubes generated with, from top to bottom 150, 100, 75, 50, 25, 15, 10, 6, 4, 2, and 1 channels respectively. For reference we plotted the predictions
in LP00: in dashed lines are the predictions for thin channels, in solid lines for very thick channels.
thin regime; unless, N < ∆v
this behaviour.
σL ∼ 5 − 6. Figure 2 is consistent with
N ≈
∆v
σL (cid:18) L
r (cid:19)m/2
(11)
In Figure 2 we can see that the power spectrum has an addi-
tional deviation from the power-law at small scales. For that rea-
son, to measure numerical values of the spectral indices, we have
restricted ourselves to the wave-number range from log k ∼ 1.3 to
2.2, where the power spectra in Fig. 2 is close to a power-law. The
slopes (spectral indices) of the power spectra obtained with a linear
least-squares method are shown in Table 2.
From Fig. 2 (and Table 2) we can see, indeed, a steepening
of the power-spectrum as we increase the thickness of the velocity
slices. However, the spectral-indices in thin channels do not appear
to reach an asymptotic value, instead they keep getting shallower
as the velocity resolution increases. This behaviour arise due to the
limited number of emitters along the LOS: with only 216 emitters
along each LOS, to distribute in a large number of velocity chan-
nels, there will be many channels empty. This produces artificial
sharp edges, and introduce high frequency components. Rendering
in a shallower spectrum at large wave-numbers. To avoid this prob-
lem, in Lazarian et al. (2001) it was suggested as optimal to search
for thin slice asymptotics in the channels that are just thin enough,
at the smallest scale r of interest. The expression in eq. (11) is
somewhat tricky, because we need to know the value of m, to get
the number of channels to be used to best determine m! But, it al-
lows to check with emulated observations if the predictions hold.
Our results, in Table 2, are in agreement with the prediction in
Lazarian et al. (2001) that more channels are required when steeper
velocity fields are studied. For real observations the problem is al-
leviated because in that case we are not limited by the number of
emitters and we can use the maximum velocity resolution with-
out having empty channels. In Lazarian et al. (2001) the factor of
∆v/σL in eq. (11) was estimated to be ∼ 2.5, in the assumption of
Gaussian statistics. In practice we can use the standard deviation of
the velocity in the LOS velocity cube for σL, and ∆v is the range
of velocity covered in our PPV cubes. This is important to keep in
mind because a side effect of the modification of the spectral in-
dex is an appreciable change in the velocity range of the velocity
cubes. To get the factor (L/r) we need to take the maximum wave-
number used to obtain the power-spectra. For the results presented
in Table 2 it was k ∼ 102.2, so L/r = k/2π ∼ 25. This way
we generated PPV cubes with the velocity resolution according to
6
Esquivel et al.
Table 2. Measured emissivity spectral indices of Fig. 2
Density index n = −2.5
Density index n = −11/3
Number
of
velocity
channels
Velocity index (µ)
-3.3
-3.67
-4.0
-3.3
-3.67
Predicted spectral index (γ)
-2.35
-2.17
-2.00
-2.67
-2.6
150
100
75
50
25
15
10
6
4
2
1
-2.02
-2.05
-2.07
-2.11
-2.22
-2.34∗
-2.44
-2.47
-2.47
-2.46
-2.49
-1.97
-2.00
-2.02
-2.06
-2.18∗
-2.33
-2.42
-2.52
-2.53
-2.52
-2.49
-1.93
-1.95
-1.97
-2.02∗
-2.15
-2.33
-2.46
-2.56
-2.55
-2.54
-2.49
-2.25
-2.27
-2.30
-2.35
-2.54
-2.74∗
-3.01
-3.13
-3.13
-3.13
-3.65
-2.19
-2.22
-2.24
-2.29
-2.47
-2.74∗
-2.99
-3.23
-3.29
-3.30
-3.65
-4.0
-2.5
-2.15
-2.18
-2.20
-2.25
-2.44∗
-2.74
-3.02
-3.39
-3.44
-3.45
-3.65
∗ With this velocity resolution (number of channels) we get the better
match for the analytical predictions in thin channels.
Table 3. Emissivity spectral indices in barely thin chanels
Density
index
Velocity
index
Num. of
channels
-2.5
-2.5
-2.5
-3.67
-3.67
-3.67
-3.3
-3.67
-4.0
-3.3
-3.67
-4.0
35
42
54
35
42
54
Predicted Measured
γ
-2.35
-2.17
-2.00
-2.85
-2.67
-2.50
γ
-2.2
-2.1
-2.0
-2.4
-2.3
-2.2
eq. (11), and measure the power spectrum in velocity channels, in
Table 3 we display the results. From Table 3 we can see that the
measured spectral indices within 15% error of the predictions, but
always shallower than expected. This is an indication, that the noise
arising from the limited number of emitters is present even in chan-
nels barely thin, although not near as large as that present for PPV
cubes constructed with more channels (see Table 2).
Could deviations above be due to correlations between the
density and the velocity fields? The answer is no: to illustrate this
point, we can compare the power spectrum of the MHD simulations
to that of an ensemble of synthetic PPV cubes with the same spec-
tral indices but no velocity-density correlations whatsoever. For this
comparison to work we must use PPV cubes constructed from data
with the same power-law inertial range. For that reason, we mod-
ified the spectral index of the density field from the simulations
to −11/3 (close to the original index, so the modification on the
velocity-density correlations is minimal). With this density field,
we then constructed PPV cubes using the modified velocity output
from the simulations and 60 synthetic velocity fields. All the ve-
locity cubes used had the same spectral index (also −11/3). The
power spectra in velocity channels is shown in Fig. 3. In this fig-
ure we can appreciate that apart of some 'wiggles' at small values
of k, due to lack of enough statistics (see discussion below) the
power-spectra is practically identical whether the density field is
correlated with the velocity field or not.
Figure 3. Power spectrum in velocity channels for PPV produced with den-
sity and velocity cubes from MHD simulations -- with spectral indices mod-
ified to −11/3 -- (a), and with the same density as the upper panel but an
ensemble of Gaussian cubes with the same −11/3 index (b). The dotted
lines correspond to the average spectrum in velocity channels for PPV with
different number of channels, in both panels from these correspond from
top to bottom to 150, 100, 75, 50, 25, 15, 10, 6, 4, 2, and 1 channels re-
spectively. We see larger departures from power-law behaviour at lower k in
(a) because in (b) we averaged over an ensemble, see discussion on the text.
Also, in both cases for reference we plotted the predictions of the power
spectrum in thick channels (solid lines), and in thin channels (dashed lines).
3.4 More on the numerical limitations
As we saw in the previous section, one important practical problem
with the determination of the spectral indices through VCA is to
measure with accuracy the slopes (precisely the spectral indices) in
the plots. We can notice from each plot in Fig. 2, two departures
from well behaved power-laws. At the left end of the spectrum
(low k numbers, large spatial scales) we have some random wig-
gling, while at the right end of the curves we see a systematic flat-
tening of the power spectrum. The latest is increasingly evident for
PPV cubes formed with larger number of channels (with thinner ve-
locity channels), and was referred as 'shot-noise' in Lazarian et al.
(2001).
3.4.1 Departures from power-law at low wave-numbers
In the spirit of avoiding other sources of uncertainty, we will use
simple Gaussian fields with prescribed power-law spectrum as ve-
locity fields, mapped with a constant density field to produce PPV
Velocity statistics from spectral line data
7
Figure 4. Power spectra in velocity channels for an ensemble of 60 Gaussian cubes of velocity with index −11/3 and constant density. The dotted lines are
the average of the 2D power spectra in velocity channels for individual PPV cubes constructed with the number of channels as denoted in the upper right
corner in each plot The solid lines correspond to the ensemble average, for visual purposes they are shifted vertically by 0.5.
cubes. We generated an ensemble of 60 Gaussian cubes with a pre-
scribed spectral index of −11/3. Then we applied to them the
VCA, that is, we obtained PPV cubes with different number of
channels (different velocity resolution) and compute the 2D power
spectra in velocity channels. The results are shown in Fig. 4. From
the figure is evident that these departures from strict power-laws
at low k numbers are due to the lack of sufficient statistics at the
largest scales. Power spectra in velocity channels for individual ve-
locity cubes differ very little from intermediate to large k numbers,
but differ significantly at small k, showing some 'wiggles'. How-
ever, the ensemble average converges to a power-law that extends
from the largest spatial scales to intermediate spatial scales, until
a systematic flattening of the power spectrum at large k becomes
evident. This problem may ocurr in actual observations, where the
low wave-numbers (large scales) are affected by the general size
and shape of the object under study.
3.4.2 Departures from power-law at high wave-numbers
One of the most important limitations that we to construct spec-
tral lines with numerics, is the finite (and very small) number of
emitters in a particular direction; in our simulations we have 216
density elements in each LOS to distribute in velocity channels. If
we impose a high velocity resolution in the PPV cubes (i. e. large
number of channels), the spectral lines instead of being smooth
will have sharp edges and empty channels. This artificial jumps
introduce high frequency components in the spectrum and there-
fore an increase of the power spectrum at large wave-numbers. In
Lazarian et al. (2001) this was regarded as 'shot-noise'. This is true
in the sense that the flattening becomes larger as we increase the
velocity resolution. But is more complicated that the usual shot-
noise, that can be easily described in terms of Poisson statistics.
This flattening of the spectrum is particularly problematic because
it increases as we go into thinner channels, and the predictions in
LP00 to extract the velocity spectral index were done precisely in
an asymptotic regime for thin channels.
A natural way to avoid the problem of this flatening of the
power spectrum is simply to increase the number of emitters along
the LOS. This can be accomplished increasing the size of the data
cubes, which translates into more computing power required to
analyse the power-spectrum (and more importantly more comput-
ing power to generate larger MHD cubes). With the resources at our
hand, we increased the number of emitters along each LOS by cre-
ating an ensemble of 50 2D Gaussian velocity fields of size 15362.
All with spectral index of −8/3, Kolmogorov in 2D. Analogously
to the 3D ensemble we simulated observations and analysed the
8
Esquivel et al.
Figure 5. Comparison of the power spectrum in velocity channels for an
ensemble of 60 velocity cubes (a) with an ensemble of 50 squares (b). The
cubes are 2163 in size and the squares are 15362. All velocity fields have
Kolmogorov spectral indices (−8/3 for 2D and −11/3 for 3D), and the
PPV were obtained with a constant density. The dashed line on the lower
panel is to compare the inertial range achievable in both cases.
power spectrum in velocity slices. In this case instead of PPV cubes
we have simply 'PV arrays', and the power spectrum computed in
slices is one-dimensional. In Fig. 5 we compare the ensemble av-
erage power-spectrum in velocity channels for both, 3D and 2D
cases. It's clear the increase in wave-number range acquired with
the 2D data, in which we have about half decade more. The relevant
result of this comparison is to show that the flattening of the power
spectrum for the thinnest channels is present in both cases, at large
k. But for the 3D case this flattening is present earlier than in the
2D case.
It's important to stress that this is NOT a real dynamical ef-
fect, the cubes used to construct the PPV cubes are just Gaussian
velocity and constant density.
Another way to alleviate the problem of limited number of
emitters is to include an intrinsic width to each emitting element.
We acomplished this task by adding some Gaussian smoothing on
each spectral line (i. e. convolving a Gaussian profile along each
LOS). This is somewhat equivalent to introduce a thermal width
to the emitters3. Some examples of the spectral lines in our PPV
cubes, with and without an intrinsic width are shown in Figure 6.
Figure 6. Examples of spectral lines in one of our PPV cubes with differ-
ent resolutions (the number of channels is indicated in each panel.). On the
left we have spectra from PPV cubes with no width associated to the emit-
ting elements, on the right the same spectra after adding a significant width
(Gaussian with FWHM of ∆/60). All the velocity axis contain the entire
velocity range in the PPV, and the intensities are normalized for visual pur-
poses.
Introducing an intrinsic width to the PPV cubes makes the spectral
lines to look more realistic because it smooths sharp edges due to
the limited numerical resolution. However, it also limits the reso-
lution that can be used to look for thin channel asymptotics. The
criterion in LP00 for thin or thick channels, was based on an 'ef-
fective width' of the velocity channels, which included not only the
thickness of the channel but also a thermal width vT as:
δvef f ≈ (cid:0)δv2 + 2v2
T(cid:1)1/2
.
(12)
The additional intrinsic width limits the resolution attainable
in two ways. For a fixed maximum k scale the velocity channels
need to be narrower to remain in the thin regime, with a minimum
number of channels given by:
N & (cid:20) σ2
∆v2 (cid:16) r
L(cid:17)m
L
v2
T
∆v2(cid:21)−1/2
.
− 2
(13)
And for a fixed velocity resolution it reduces the wave-number up
to which the channel remains thin to
k . 2π(cid:20) 1
σ2
L (cid:18) ∆v2
N 2 + 2v2
T(cid:19)(cid:21)−1/m
.
(14)
The criteria in eq.(14) is more restrictive, beyond that scale the ve-
locity channels are entering in a transition to the thick regime. The
condition in eq.(13) is less restrictive, in the sense that any slice
with that minimum number of channels is to be considered thin.
However, there is no reason to slice narrower than that, beacuse the
thermal width smears velocity fluctuations on smaller scales. If we
go to yet thinner channels we reduce the amplitude of the signal,
but do not see any further change of spectral index. In Figure 7 we
show the power spectra in velocity channels for PPV cubes where
we included an intrinsic width similar to the width of the barely
thin channels (thin down to the scale corresponding to the limit
we chose previously to measure the power spectra, log k ∼ 2.2).
From the figure we can notice immediately that the rising tail at
3 We used this width as free parameter, therefore we rather use the term
'intrinsic' instead of 'thermal' width, wich stricltly speaking, is determined
by physical mechanisms beyond the scope of our simulations.
Velocity statistics from spectral line data
9
Table 4. Spectral indices in velocity channels with shear
Num. of
channels
No
shear
Shear magnitude (in units of σvz /L)
1
20
3
5
10
150
100
75
50
25
15
10
6
4
2
1
-2.2
-2.3
-2.3
-2.4
-2.6
-2.9
-3.2
-3.4
-3.4
-3.4
-3.7
-2.3
-2.3
-2.4
-2.4
-2.7
-3.0
-3.1
-3.3
-3.3
-3.3
-3.7
-2.4
-2.4
-2.5
-2.6
-2.8
-3.0
-3.1
-3.1
-3.1
-3.1
-3.7
-2.4
-2.5
-2.5
-2.6
-2.9
-3.1
-3.3
-3.1
-3.1
-3.1
-3.7
-2.7
-2.7
-2.8
-2.9
-3.2
-3.4
-3.4
-3.3
-3.3
-3.2
-3.7
-3.1
-3.1
-3.2
-3.3
-3.7
-4.1
-4.0
-3.7
-3.5
-3.4
-3.7
z axis and added a component to the velocity cube (the vz cube)
following
vz(x, y, z) = v0z(x, y, z) + Cσvz
x
L
z,
(15)
where vz is the new velocity cube used to generate the PPV cubes,
v0z is the original velocity cube before introducing the shear, σvz is
the velocity dispersion of v0z, L is the cube size in the spatial coor-
dinates (x, y), and C is a parameter we introduce to vary the mag-
nitude of the shear introduced. The shear introduced is the same on
all scales, and equal in magnitude, to Cσvz /L. We generated PPV
cubes adding a linear shear with magnitude up to 1, 3, 5, 10, and
20 times maximum velocity dispersion over the largest scale (i. e.
with a velocity component with C = 1, 3, 5, 10, and 20 in eq.
15). As before, we measure the spectral indices in the range less
affected by numerical noise, log k from ∼ 1.3 to ∼ 2.2. The range
used is somewhat arbitrary, but the purpose of the study at this point
is just comparison with the case of no shear. The measured slopes
are summarized in Table 4. The introduced shear is the same on all
scales, while the turbulent shear scales as ∼ k2/(m−2). The mag-
nitude of the external shear is the same as the magnitude of the tur-
bulent shear when k ≈ 2πC 3/2. Therefore the external shear con-
tributed initially to the power spectrum only at the largest scales,
but eventually became important at intermediate to small scales as
its magnitude increased (see Fig. 8). Our results show that the effect
of an external shear becomes important only when it is comparable
to the local turbulent shear. In our example that was when the max-
imum shear reached about 5 times the velocity dispersion over the
largest scale. The relative differences of the spectral indices mea-
sured compared with the case of no shear are below 10%, for the
cases of less shear (C . 5). It is noticeable that the larger distor-
tion happens at lower k, even when its magnitude is comparable to
the turbulent shear at smaller scales. This is expected, given that the
shear is regular, and represents a large scale motion. And holds if
we measure the slope of the power spectrum in intermediate spatial
scales, which are available using synthetic data. With actual obser-
vations, we can go further into smaller scales and avoid the effects
of a stronger shear. The shear due to turbulence dominates in the
galaxy, and its the effect to the VCA is marginal.
5 ANISOTROPIC TURBULENT CASCADE
Magnetic field plays a crucial role in the dynamics of inter-
stellar turbulence, and makes the turbulent cascade anisotropic
(Montgomery 1982; Higdon 1984). In a turbulent magnetized me-
Figure 7. Same as Fig. 2 but including an intrinsic width to the emitters.
Each emitting element is a Gaussian with a FWHM of ∆v/60. We can see
how the flatening of the power spectrum at large k decreases dramatically.
high wave-numbers dramatically decreases. In the cases where the
velocity fluctuations at small scale structure are relativelly small
(in Fig. 7, µ = −4) the transition from thin to thick channels as
we go into smaller scales (large k) is evident. This corresponds to
predictions in LP00. We computed the spectral indices in velocity
channels, for our previous inertial range with slices of thickness
according to eq.(13), and also for the conservative inertial range
from eq.(14) for the same velocity resolutions as before. In both
cases we observe that the measured spectral indices doesn't change
significantly from the earlier estimates with barely thin channels
(presented in Table 3). This is understandable because the intrin-
sic width we included alleviates the problem of small number of
emitters only on scales smaller than such width.
This shot-like noise presents an obstacle if we want to obtain
the velocity index of data cubes generated synthetically. However,
we must stress that the problem is not expected to be present in
real observations, where the number of emitters along the LOS is
much larger. This was confirmed through observations of the Small
Magellanic Cloud (Stanimirovi´c & Lazarian 2001). The spectral
indices there, measured in thinner and thinner channels do tend to
an asymptotic value, not being affected by 'shot-like noise' allowed
to measure accurately the power-spectrum in velocity channels.
4 EFFECT OF THE SHEAR ON THE VCA
The VCA is sensitive to density and velocity fluctuations at a partic-
ular scale. On the other hand, shear as any organized motion, like
galactic rotation, would correspond to the smaller wave-numbers
(the largest scales, sometimes much larger that the largest scale in a
PPV cube). If we choose the right range in k to perform the VCA,
we should avoid any interference of this ordered motions with the
analysis. To explore numerically the effect of a linear shear on the
VCA we used modified data cubes of spectral indices of −11/3 for
both velocity and density fields. We chose the LOS to be along the
10
Esquivel et al.
Figure 8. Power spectra in velocity channels including linear shear. Each
panel correspond to a different magnitude of the maximum shear, as labelled
in the title of each plot. The different lines in each panel correspond from
top to bottom to the average power spectrum in velocity channels for PPV
cubes constructed with 150, 100, 75, 50, 25, 15, 10, 6, 4, 2, and 1 channels
respectively. The small arrows show at which wave-number the magnitude
of the external shear equals the magnitude of the turbulent shear.
dia, the energy on the large scale motions is larger than that on the
small scale structures; however the local magnetic field strength
remains almost the same. Thus, is easier to bent magnetic field
lines on the large scale by turbulent motions, but on the small
scale becomes more difficult. Another way to think about this
phenomenon is noting that hydrodynamic motions can easily mix
magnetic field lines on the small scale, rather than bent them, in
the direction perpendicular to the mean magnetic field (see a dis-
cussion in Lazarian & Vishniac 1999). As a result we get elon-
gated eddies, relative to the magnetic field that become even more
elongated as we go further into smaller scales. A model for in-
compressible anisotropic turbulence is that of Goldreich & Sridhar
(1995). This model has been supported by numerical simu-
lations (e.g. Cho & Vishniac 2000; Maron & Goldreich 2001;
Cho, Lazarian & Vishniac 2002). Compressible anisotropic MHD
turbulence has been studied in Lithwick & Goldreich (2001) and
Cho & Lazarian (2002, 2003).
5.1 Magnetic Field Direction: Anisotropy in Correlation
Functions
In an isotropic turbulent cascade correlations depend only on the
separation between points. Contours of equal correlation are circu-
lar in that case. If on the other hand the turbulence is anisotropic,
two point statistics are also anisotropic. In this case contours of
iso-correlation become elongated, with symmetry axis given by the
magnetic field direction. The technique presented here is similar to
Figure 9. Contours equal correlation of 2D maps obtained with the PPV
data. (a): contour map of the correlation of emissivity at the centroid of
the lines. (b): contours of correlation in the integrated emissivity. (c), (d):
contours in two individual channel maps. These were obtained in a PPV
cube with 25 channels (the velocity range is from Vmin = −1.27 to
Vmax = 1.451, and consequently the channel thickness is δV = 0.109).
(c) corresponds to a velocity centred at Vz = −1.11, while the (d) to a
velocity of Vz = 0.741. All the distances in the figures are in grid units.
one proposed to study magnetic field direction using synchrotron
maps (Lazarian 1992). Iso-contours of correlation functions can al-
low us to determine the direction of the magnetic field. We calculate
two point correlation for different 2D maps generated with our data.
The correlation function between the two scalar functions f (x) and
g(x) can be computed from
C(r) = h(f (x) − hf (x)i) · (g(x + r) − hg(x)i)i
(16)
σf σg
,
where r is the separation between two points or 'lag', x is the spa-
tial separation, σf and σg are the standard deviations of f and g
respectively. Here the average is performed on all the radial com-
ponents and azimuthal angles.
With the original cubes from simulations (no artificial modi-
fication of the spectral indices), we calculated the auto-correlation
(correlation in eq. (16) with f (X ) = g(X)) on various 2D maps:
integrated intensity (integrated along the LOS, equivalent to a 2D
map of column density); emissivity at the centroid of the lines, ob-
tained as:
jc(X) = R j(X )vdv
R j(X )dv
,
(17)
where the emissivity j(X ) in our case is proportional to the first
power of the density; and maps of emissivity in individual veloc-
ity channels. Contour maps of iso-correlations calculated of these
maps are given in Fig. 9. We can see that anisotropy is present in
the contours plotted. It is evident in the case of the emissivity at the
centroids of the lines, as well as in the integrated emissivity. In in-
dividual velocity slices, we found that the majority of the velocity
channels show this anisotropy in the expected direction (elongation
of the contours parallel to the mean B field), but in some slices
we see the elongation of correlation contours tilted from the mean
magnetic field direction. For Fig. 9 we choose a representative slice
to illustrate this misleading behaviour.
The main purpose of this subsection was to show an alternative
method to obtain the direction of the magnetic field, and also to
illustrate that anisotropic cascade is present in our simulations.
The anisotropy based technique could be a unique tool for
magnetic field studies, especially when other techniques, e.g. based
on dust alignment fail (see Lazarian 2000 for a review on grain
alignment).
5.2 Anisotropic cascade & VCA.
It's important to test the VCA, which was developed considering
isotropic turbulence, in a more realistic case where the magnetic
field determines a preferential direction of the turbulent motions.
With that purpose we constructed PPV cubes choosing the LOS to
be parallel to the mean magnetic field (along the x axis) and com-
pare the results with the case where the magnetic field is perpendic-
ular to the LOS (z axis). For this exercise we use a data cubes cor-
rected to Kolmogorov spectral indices (−11/3 in 3D). The power
spectrum in velocity channels is shown in Fig. 10. In the figure we
can see that the slopes of the spectrum (spectral index) are almost
identical whether the observer is looking parallel or perpendicular
to the mean magnetic field. We performed a least-squares linear fit-
ting and found the differences of the slopes in every case are well
within the fitting uncertainties.
We conclude that VCA is applicable to anisotropic turbulence
studies.
6 SUMMARY AND CONCLUSIONS
In this paper we have tested the VCA technique numerically. We
used compressible MHD simulations, and an ensemble of Gaus-
sian fields to emulate spectroscopic observations (PPV data cubes).
We computed the power spectra in velocity channels to test the an-
alytical predictions in LP00.
To obtain the PPV data we modified the output of the sim-
ulations to force power-law indices to obtain density fields with
spectral indices of −2.5 and −11/3, and velocity fields with spec-
tral indices of −3.3 to −4.0). The spectral indices derived from our
emulated observations are within a 15% error of the analytical pre-
dictions. We confirmed trends predicted in LP00, like the steepen-
ing of the emissivity index as we increase the width of the velocity
channels, or that the velocity fluctuations dominate the emissivity
for thin velocity channels. This last result is in agreement with a
previous study by Pichardo et al. (2000), and is important because
warns observers against interpret any structure in PPV as density
enhancements or 'clouds' (see also discussion in LP00).
We showed that the deviations from power-law behaviour at
the largest spatial scales (small k numbers) are due to lack of good
statistics at those scales. This was comparing the results obtained
with PPV produced with a constant density and the modified veloc-
ity field of spectral index −11/3, to those using an ensemble of 60
Gaussian cubes with the same index.
An important source of error addressed is a flattening of the
power spectrum at large wave-numbers. This is due to the lim-
ited number of emitters along the LOS. We illustrated this point
comparing the results obtained with the ensemble of 60 cubes with
216 emitters along the LOS, and an ensemble (two-deimensional)
with 1536 emitters. We also try including an intrinsic width to the
Velocity statistics from spectral line data
11
emitters along the LOS to smooth the sharp edges arising from
the limited number of emitters. This procedure reduced the flat-
tening of the spectrum at large k numbers. But, also increases the
effective width of the channels, reducing our ability to make use of
the scales on wich we reduced such flattening of the power spec-
trum. This problem was found to be the most restrictive to our
work. However, is exclusive of numerical simulations, and is not
expected to be important in real observations (like those presented
in Stanimirovi´c & Lazarian 2001), where the number of emitting
elements is essentially infinite.
We introduced linear shear to our data, and showed that its
effect only becomes important when the magnitude of the shear is
comparable to the local turbulent shear. We showed, that the effect
of large scale shear is marginal at small scales. And that avoid-
ing the largest scales to computed the power-spectrum allowed to
recover the turbulence statistics. This shows that for ordinary ob-
servational situations we can safely neglect shear in the VCA.
We analysed the cross-correlation for 2D maps of integrated
intensity and emissivity at the centroids of the lines, and showed
that they are anisotropic (elongated), with symmetry axis defined
in the direction of the mean magnetic field. This constitutes a
new technique to study the direction of the magnetic field. We
hope that the anisotropies will reveal the magnetic field within
dark clouds, where grain alignment and therefore polarimetry fails.
From the point of view of VCA, we illustrated that the effect of the
anisotropy of the turbulence is marginal.
ACKNOWLEDGMENTS
We thank the referee for valuable suggestions that improved this
work. AE acknowledges financial support from CONACyT (Mex-
ico). Research by AL and JC is supported by NSF grant AST 01-
25544.
REFERENCES
Bond J. R., Kofman L., Pogosyan D., 1996, Nature, 380, 603
Brunt C. M., Heyer M. H., 2002a, ApJ, 566, 276
Brunt C. M., Heyer M. H., 2002b, ApJ, 566, 289
Cho J., Lazarian A., 2003, in Rast M., ed., Acoustic emission and
scattering by turbulent flows, Springer Lecture Notes in Physics,
in press
Cho J., Lazarian A., 2002, Phys. Rev. Lett., 88, 245001
Cho J., Lazarian A., Vishniac E. T., 2003, Passot T., Falgarone E.,
eds, Simulations of magnetohydrodynamic turbulence in astro-
physics, Springer Lecture Notes in Physics,in press
Cho J., Lazarian A., Vishniac E. T., 2002, ApJ, 564, 291
Cho J., Vishniac E. T., 2000, ApJ, 539, 273
Goldreich P., Sridhar S., 1995, ApJ, 438, 763
Gott J. R., Mellot A. L., Dickinson M., 1986, ApJ, 306, 341
Heyer M. H., Schloerb F. P., 1997, ApJ, 475, 173
Higdon J. C., 1984, ApJ, 285, 109
Jiang G., Wu C. 1999, J. Comp. Phys., 150, 561
Lazarian A., 1992, Astron. and Astrophys. Transactions, 3, 33
Lazarian A., 1995, A&A, 293, 507
Lazarian A., 1999, Plasma Turbulence and Energetic Particles in
Astrophysics, Proceedings of the International Conference, Cra-
cow (Poland), 5-10 September 1999, Eds.: Michał Ostrowski,
Reinhard Schlickeiser, Obserwatorium Astronomiczne, Uniwer-
sytet Jagiello´nski, Krak´ow 1999, p. 28-47., 28
12
Esquivel et al.
Figure 10. Power spectra in velocity channels for observations with the observer at two different positions. Each panel correspond to the average spectrum
in PPV cubes with different number of velocity channels (the number of channels is indicated in the upper right corner on each). The solid lines correspond
to the LOS parallel to the mean magnetic field; for the dotted lines the LOS is perpendicular to the mean magnetic field. All the PPV cubes were constructed
from the same set of density and velocity data, 'modified' to 3D spectral indices of −11/3.
Spangler S. R., Gwinn C. R., 1990, ApJL, 353, L29
Scalo J. M., 1987, ASSL Vol. 134: Interstellar Processes, 349
Stanimirovi´c S., Lazarian A., 2001, ApJL, 551, 53
Scoccimarro R., Couchman H. M. P., Frieman J. A., 1999, ApJ,
517, 531
Lazarian A., 2000, ASP Conf. Ser. 215: Cosmic Evolution and
Galaxy Formation: Structure, Interactions, and Feedback, 69
Lazarian A., Pogosyan D., 2000, ApJ, 537, 72
Lazarian A., Pogosyan D., Esquivel A., 2002, in Taylor R., Lan-
decker T. L., Willis A. G., eds, ASP Conf. Ser. Vol. 276, Seeing
through the dust. Astron. Soc. Pac., San Francisco, p. 182
Lazarian A., Pogosyan D., V´azquez-Semadeni E., Pichardo B.,
2001, ApJ, 555, 130
Lazarian A., Vishniac E. T., 1999, ApJ, 517, 700
Lithwick Y., Goldreich P., 2001, ApJ, 562, 279
Liu X., Osher S., 1998, J. Comp. Phys., 141, 1
Maron J., Goldreich P., 2001, ApJ, 554, 1175
Mellot A. L., Cohen A. P., Hamilton A. J. S., Gott J. R., Weinberg
D. H., 1989, ApJ, 345, 618
Miesch M. S., Bally J., 1994, ApJ, 429, 645
Monin A. S., Yaglom A. M., Statistical fluid dynamics: mechanics
of turbulence, Vol. 2., MIT press, Cambridge, MA
Montgomery D., 1982, Phys. Scripta, 2, 83
Narayan R., Goodman J., 1989, MNRAS, 238, 963
Pichardo B., V´azquez-Semadeni E., Gazol A., Passot, T.,
Ballesteros-Paredes, J., 2000, ApJ, 532, 353
Rosolowsky E. W., Goodman A. A., Wilner D. J., Williams J. P.,
1999, ApJ, 524, 887
|
astro-ph/0311542 | 1 | 0311 | 2003-11-24T18:17:27 | How old are the HII Galaxies? | [
"astro-ph"
] | Using a novel approach we have reanalized the question of whether the extreme star forming galaxies known as HII galaxies are truly young or rejuvenated old systems. We first present a method of inversion that applies to any monotonic function of time describing the evolution of independent events. We show that, apart from a normalization constant, the ``true'' time dependence can be recovered from the inversion of its probability density function. We applied the inversion method to the observed equivalent width of Hbeta (EW(Hbeta)) distribution for objects in the Terlevich and collaborators Spectrophotometric Catalogue of HII galaxies and found that their global history of star formation behaves much closer to the expectations of a continuos star formation model than to an instantaneous one. On the other hand, when the inversion method is applied to samples within a restricted metallicity range we find that their history of star formation behaves much closer to what the instantaneous model predicts.
Our main conclusion is that, globally, the evolution of HII galaxies seems consistent with a succession of short starbursts separated by quiescent periods and that, while the emission lines trace the properties of the present burst, the underlying stellar continuum traces the whole star formation history of the galaxy. Thus, observables like the EW(Hbeta) that combine an emission line flux, i.e. a parameter pertaining to the present burst, with the continuum flux, i.e. a parameter that traces the whole history of star formation, should not be used alone to characterize the present burst. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 8 (2003)
Printed 26 October 2018
(MN LATEX style file v1.4)
How old are the HII Galaxies?
Roberto Terlevich1⋆, Sergiy Silich1, Daniel Rosa -- Gonz´alez2 and Elena Terlevich1
1 INAOE, Luis Enrique Erro 1. Tonantzintla, Puebla 72840. M´exico.
2 Astrophysics Group, Blackett Laboratory, Imperial College, Prince Consort Road, London SW7 2BW.
Accepted . Received ; in original form 26 October 2018 VERSION: v012
ABSTRACT
Using a novel approach we have reanalized the question of whether the extreme
star forming galaxies known as HII galaxies are truly young or rejuvenated old systems.
We first present a method of inversion that applies to any monotonic function of
time describing the evolution of independent events. We show that, apart from a nor-
malization constant, the "true" time dependence can be recovered from the inversion
of its probability density function.
We applied the inversion method to the observed equivalent width of Hβ
(EW(Hβ)) distribution for objects in the Terlevich and collaborators Spectrophoto-
metric Catalogue of HII galaxies and found that their global history of star formation
behaves much closer to the expectations of a continuos star formation model than to an
instantaneous one. On the other hand, when the inversion method is applied to sam-
ples within a restricted metallicity range we find that their history of star formation
behaves much closer to what the instantaneous model predicts.
Our main conclusion is that, globally, the evolution of HII galaxies seems con-
sistent with a succession of short starbursts separated by quiescent periods and that,
while the emission lines trace the properties of the present burst, the underlying stellar
continuum traces the whole star formation history of the galaxy. Thus, observables
like the EW(Hβ) that combine an emission line flux, i.e. a parameter pertaining to the
present burst, with the continuum flux, i.e. a parameter that traces the whole history
of star formation, should not be used alone to characterize the present burst.
Key words:
1
INTRODUCTION
HII galaxies are dwarf emission line galaxies undergoing a
burst of star formation. They are characterized by strong
and narrow emission lines originated in a giant star forming
region which dominate their observable properties at optical
wavelengths. Most HII galaxies are in fact Blue Compact
Galaxies (BCG's), but due to the different selection criteria
only a small percentage of BCG's, i.e. those with the largest
emission line equivalent widths, are HII galaxies. We will
stick to the name "HII galaxies " to refer to the star forming
systems selected from objective prism surveys and having
strong narrow emission lines.
Various studies of the spectroscopic properties of HII
galaxies in optical wavelengths revealed systems of very low
heavy element abundances and high rates of star formation.
Earlier morphological studies have suggested that a large
⋆ RT and ET, Visiting Fellows at the Institute of Astronomy,
Cambridge, U.K.
c(cid:13) 2003 RAS
proportion of the sample of HII galaxies observed are com-
pact and isolated (Melnick 1987). This, together with the
spectroscopic properties indicating low to very low metal-
licity plus a very young stellar content, led workers in the
field ever since their discovery, to question whether these
systems are truly young galaxies or made up by a few
bursts separated by long quiescent periods in the lifetime of
the galaxy. Reviews of the general properties of HII galax-
ies can be found in Melnick (1987), Terlevich (1988), the
"Spectrophotometric Catalogue of HII Galaxies " (hereafter
SCHG, Terlevich et al. 1991), Stasi´nska & Leitherer (1996),
Telles & Terlevich (1993; 1995; 1997), Telles, Melnick & Ter-
levich (1997), and more recently in the excellent review by
Kunth & Ostlin (2000).
More than 20 years ago Dottori (1981) suggested the use
of the EW(Hβ) to detect differences in age among HII re-
gions. Dottori & Bica (1981) applied the EW(Hβ) method to
31 LMC and SMC HII regions and found an uneven distribu-
tion of ages with most HII regions clustering at an EW(Hβ)
of about 70A and only 5 HII regions with EW(Hβ) >120 A.
2
Roberto Terlevich et al.
This result led them to suggest that a galaxy-wide burst of
star formation did occur 6 to 7 Myr ago in the Magellanic
clouds.
The SCHG compiles, among other parameters, line ra-
tios and equivalent widths for several hundred HII galaxies,
and gives the opportunity of investigating the ages of ac-
tively star forming galaxies to find whether there is among
them a truly young system. The SCHG is particularly ap-
propiate for this task because, due to the selection criterion
used to find them, the HII galaxies in the SCHG are proba-
bly the youngest systems that can be studied in any detail.
This is due to the fact that the SCHG samples the narrow
emission line galaxies with the highest equivalent width in
their emission lines, and is particularly biased in the local
universe (at z<0.04) towards strong and compact emission in
[OIII] 5007A. This bias is introduced by the technique used
in all the 3 sources of data, the Cambridge, Tololo and Uni-
versity of Michigan surveys, that searched for strong emis-
sion using a Schmidt camera (either the UK or the Tololo
Schmidt at the AAT and CTIO respectively) equipped with
low dispersion objective prisms and IIIaJ emulsion, a com-
bination that produces a sensitivity range from 3500A to
5300A.
One of the early surprising results of the SCHG was
that even in this strong emission line biased sample, there
are less than 10% of the systems with EW(Hβ) >150 A
(Fig. 1, left). In contrast, as shown in Fig. 1, right, a quick
analysis of the predictions for a single burst indicates that if
star formation proceeeds uniformly with time, about half of
the regions of star formation should have an EW(Hβ) larger
than about 150 - 200 A depending on model assumptions (all
our synthesis computations use the SB99 code of Leitherer
and collaborators (1999)).
Many explanations have been put forward for the lack
of systems with high EW(Hβ). Chiefly among them, time
dependent escape of ionizing photons, dust affecting either
the ionizing radiation or the visibility occurring preferen-
tially during the first 3 Myrs, the presence of an underlying
old population, uncertainties with the models, etc. The wide
variety of possibilities has had the negative effect of almost
halting the research in this important area.
We decided to take a fresh and different approach based
as much as possible on unsupervised analysis of samples. We
describe here a new method that, under the assumption that
the catalogue samples a population of bursting galaxies at
different ages, allows the reconstruction of the time evolution
of the burst Balmer emission lines equivalent width, from
their observed distribution.
2 THE METHOD
2.1 Brief description
In this section we present our probability density distribu-
tion inversion method. In contrast with other methods, we
need only to assume that the birth rate of starburts in the
sample is random, i.e. there is no relation in the ocurrence of
starbursts in different galaxies, and that the time evolution
of the observed parameter is monotonic.
In fact, starbursts have a variety of ages, and hence a
range of equivalent widths of the emission lines. Two limiting
cases are commonly used to describe their time evolution.
They are the coeval starburst (SB) case, which assumes that
all stars are formed simultaneously in an instantanous SB
episode, where the characteristic time τSB is short compared
to the age of the galaxy (τSB ≪ t), and the continuous star
formation case, which assumes the star formation rate to
be constant in time. The first case is widely applied for in-
dividual, moderate mass star clusters, whereas the second
one is assumed to be an average characteristic of the mas-
sive star forming systems. In both cases the evolution of the
EW(Hβ) is a monotonically declining function of time. The
continuous star formation could also be approximated as a
sequence of small "mini-bursts" localized within a rather
small region in space and separated by short time intervals
(see, e.g. Silich et al. 2002).
If the starbursts birth rate, R(t), and the evolution of
an individual starburst emission lines equivalent width are
known, then the probability distribution of starbursts with
equivalent width is given by (e.g. Scalo & Wheeler, 2001)
ρw(EW ) = −
1
N0
R[t(EW )]
dEW /dt
,
(1)
where N0 is the total number of objects in the sample, and
the function t(EW) describes the time evolution of the EW
and is defined by the star formation mode.
Assuming a constant rate of star formation across the
volume of the catalogue,
R(t) = N0/tSF = Const
one can obtain the normalized probability density as,
ρw(EW ) = −
1
tSF
dt(EW )
dEW
,
(2)
where tSF is the total evolutionary time to be considered.
For the case of a monotonic time dependence the inverse
transformation of the EW distribution into a function EW(t)
is, from the relation (2):
t(EW ) = −
i
X
imax
tSF ρi(EW ) ∆EW =
tSF
N0
imax
X
i
∆Ni,
(3)
where the limits i and imax correspond to the bins with
equivalent width EW and EWmax respectively.
Therefore, the shape of the time dependence of the
equivalent width can be recovered from the observed dis-
tribution. On the other hand, the presence of an integration
constant shows that the characteristic star formation time
scale tSF cannot be obtained from the observed EW distri-
bution alone.
2.2 Numerical simulations
We have run a number of tests on the analytical descrip-
tion above. Here we show the case for the evolution of the
EW(Hβ) based on the expected evolution of an instanta-
neous burst.
We first generated a sample of random numbers homo-
geneously distributed in time, which we associate with the
starburst birthrate. For each value the corresponding equiva-
lent width was calculated using the predicted evolution from
SB99 models. Panels a and b of figure 2 show the run of the
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
How old are the HII Galaxies?
3
Figure 1. The equivalent width distribution of HII galaxies from the SCHG is plotted on the left. Clearly, and contrary to model
expectations, there are only a few systems with EW(Hβ) > 150A. The predicted time evolution of a coeval population for a Salpeter
IMF with 100M⊙ and 0.1M⊙ upper and lower mass limit respectively is plotted on the right. For a random population with a constant
production rate and considering that after 6.5 Myr a system will no longer be considered an HII galaxy, about equal number of systems
with ages below and above 3.2Myr are expected, i.e. about equal number of systems with EW(Hβ) above and below 150-200 A.
c
a
b
d
Figure 2. Panel a shows the sample of random numbers homogeneously distributed in time, which we associate with the starburst
birthrate. Panel b shows the adopted model, same as figure 1. Panel c shows the computed distribution function of equivalent widths.
Panel d shows the inversion of panel c distribution function using equation 3. Note that the inversion closely reproduces the input model.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
4
Roberto Terlevich et al.
burst rate and the adopted evolution of the EW(Hβ) respec-
tively. Panel c shows the resulting probability density dis-
tribution function of equivalent widths. This represents the
distribution of EW(Hβ) in the homogeneously distributed
instantaneous burst model catalogue. Using equation 3 we
inverted the catalogue EW probability density distribution
function shown in panel c and the reconstructed function
is shown in panel d. As expected from the analysis above,
the inverted distribution function reproduces remarkably
well the shape of the input time evolution of the equivalent
widths.
3 APPLICATION TO HII GALAXIES
We have applied the inversion method to several samples
of HII galaxies and giant HII regions. We show here only
the results for the SCHG that represents the more extreme
case of bias towards strong line HII galaxies. The results for
other samples will be published elsewhere but the main con-
clusions of the present paper are unchanged by the extended
sample analysis.
The results of the inversion are shown with the thick line
in figure 3 while the thin lines show the model predictions for
the instantaneous burst (left panel) and for the continuous
star formation (right panel) cases. It can be seen from figure
3 that the shape of the single burst model prediction is very
different from that of the reconstructed evolution function of
the EW(Hβ). The expected evolution of the EW(Hβ) will be
different for different initial conditions in the models like the
upper limit or slope of the IMF. We found that invariably
all the predicted evolution curves have a convex shape (see
figure 6 for different IMF values) while the inversion shows
a concave shape. In what follows we will use as reference
for the model corresponding to the instantaneous burst the
prediction for Mup =120 M⊙ and Salpeter slope. The rea-
son being that the sample is strongly biased towards high
excitation systems suggesting the need for a combination of
relatively low metal content and a hot ionizing cluster to
obtain it. This implies that stars more massive than 40-50
M⊙ should be present at zero age.
On the right panel we can see that models with contin-
uous star formation are in closer agreement with the shape
of the time evolution of the EW(Hβ). This rather surpris-
ing result perhaps indicates that the presence of older stars
is affecting the continuum luminosity, consistent with ear-
lier suggestions from, e.g. Dufour et al. (1996), Garnett et
al. (1997), Legrand et al. (2000).
3.1 Continuum colour and age.
To test on the somehow unexpected result of the previous
section, we have analyzed other time dependent parameters
that could provide independent information about the global
age or evolutionary stage of HII galaxies.
The continuum colour is one of such age indicators. Par-
ticularly sensitive to early age are those colours that like
Johnson's U-B, bridge the λλ3800A to 4000A region. For
this analysis, we selected those HII galaxies with the low-
est dust reddenning correction (about 220 HII galaxies) and
compared them with SB99 estimates of the colours of an
evolving stellar population.
In figure 4 we have plotted the EW(Hβ) and the λλ
3730/5010A colour defined as the ratio of the intensities of
the adjacent continua to the [OII]λ3727A and [OIII[λ5007A
emission lines.
Figure 4 also shows in thick lines the evolution of the in-
stantaneous and the continuous star formation models. The
digits along the curves represent the age of the models in
units of Myr.
If HII galaxies were truly evolving as continuous star
forming systems, they will not depart much from the con-
tinuous star formation line. Furthermore we will expect a
symmetric distribution with respect to the continuous star
formation curve. This, however, is not the case for the sam-
ple of HII galaxies. Most of the observed values are above
and/or to the left of the continuous star formation model
predictions.
In Figure 4 we have also plotted the lines correspond-
ing to multiple burst models. The thin solid lines represent
the evolutionary sequence path for individual bursts from
a sequence of identical instantaneous starbursts, separated
by 50 Myr quiescent intervals. The lines correspond to the
bursts starting at 50, 150, 350, and 750 Myr, respectively.
The thin dashed lines show the evolution of a second burst
which occurs 900 Myr after the initial one for a range of
bursts mass ratios. The mass ratio of the second to the first
starburst changes from the right top to the left bottom lines
as 10−2, 5 × 10−3 and 10−3, respectively.
It is possible to see that both the position and the scat-
ter of the points in the EW(Hβ) vs. colour diagram, are
consistent with a population of galaxies undergoing multi-
ple bursts of star formation during their cosmological evo-
lution. It is interesting to note the almost complete absence
of points to the right of the instantaneous case suggesting
a good agreement between models and observation. Reasur-
ingly, the highest observed EW(Hβ) is also consistent with
the model predictions.
3.2 Metallicity and age
Another parameter that is expected to change in galaxies
with age is their metal content. In particular, Oxygen that
represents more than 40 percent in mass of the metals, seems
to be a good metallicity indicator given that is mostly pro-
duced in massive stars with little time delay (see, e.g. Pagel
1998).
To investigate the behaviour of the metallicity of HII
galaxies versus their EW(Hβ) we have used the compilation
of the best determinations of O/H in HII galaxies (Deni-
col´o, Terlevich & Terlevich 2002). In figure 5 we have plotted
metallicity and EW(Hβ) for the 183 star forming galaxies of
the Denicol´o et al. compilation. Figure 5 shows a clear rela-
tion with EW(Hβ) albeit with a lot of scatter. The tendency
is in the sense that high metallicity HII galaxies show lower
EW(Hβ) while low metallicity HII galaxies have invariably
high EW(Hβ).
While a detailed analysis of the relation between metal-
licity and EW(Hβ) and its time evolution, requires galactic
chemical evolution studies, outside the scope of this work
and will be dealt with in a separate paper, the inspection
of simple multiple burst models (Pilyugin & Edmunds 1996,
Pilyugin 1999), nevertheless, allows as to predict that by
selecting a sample within a narrow range of metallicities,
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
How old are the HII Galaxies?
5
Figure 3. Application of the inversion method to the SCHG. The thick line shows the inversion results using equation 3 on the observed
equivalent width distribution of HII galaxies (See left panel of Fig. 1). The thin line shows the model predictions for an instantaneous
burst (left panel) and for continuous star formation (right panel)
Figure 4. The colour vs. equivalent width plot for the models and the 217 HII galaxies from our sample. The numbers on the model
tracks represent the star cluster age (from 1Myr to 9Myr and from 1Myr to 1000Myr for the instantaneous and continuous star forming
models, respectively). The thin solid lines show the evolution of a system following a sequence of identical bursts separated by 50 Myr
time intervals. For simplicity we plotted only 4 results, i.e. those starting at 50, 150, 350 and 750 Myr. Dashed lines display evolutionary
trends for the secondary starbursts occuring 900 Myr after the initial instantaneous burst of star formation. The ratio of the secondary
to the initial starburst mass drops from the right top to the left bottom lines like 10−2, 5 × 10−3 and 10−3, respectively.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
6
Roberto Terlevich et al.
and under the assumption that in short time scales, i.e. less
than 10 Myrs, very little (if any) contamination due to the
present burst occurs (e.g. Roy and Kunth, 1995), we may be
able to construct a sample with a narrow range in its chem-
ical evolutionary age. If this is the case, the inversion of the
EW(Hβ) distribution of a sample with a narrow metallic-
ity range may recover a star formation history more closely
related to a single event.
For this test, we have selected from Denicol´o et al.
(2002), the sub-sample of 70 galaxies covering the range 7.50
< 12 +log (O/H) < 8.25 The result of the inversion of the
distribution for this sample restricted in O/H is shown in fig-
ure 6 together with the predictions from three SB99 models.
Two of the models represented in the figure, correspond to
a Salpeter slope IMF having upper mass limits of 100M⊙
and 30M⊙ respectively, the other corresponds to an upper
mass of 120M⊙ and an IMF slope of 3.0, i.e. steeper than
the Salpeter case.
Clearly, the inversion has yielded an evolution whose
shape is very close to that of the instantaneous models; fur-
thermore the inversion seems more in agreement with an
IMF with Salpeter slope and an upper mass limit intermedi-
ate between 30M⊙and 100M⊙, or with a steeper slope IMF.
The simplest interpretation consistent with our findings
is that there are two different time scales for the evolution
of HII galaxies on the metallicity - EW(Hβ) plane.
On a time scale of about 107yr after any starburst, the
evolution on the metallicity - EW(Hβ) plane proceeds ver-
tically downwards as it is associated with a rapid decrease
of the EW(Hβ) as shown, e.g. in figure (1, right) and with
basically no change in the metal content of the ionized gas.
On time scales of order of 109yr there is the secular
or cosmological evolution of the ISM metal content which
reflects on the stellar population build-up.
The superposition of these two time scales results in the
dispersion observed in figures 4 and 5.
Our findings are also consistent with the idea that the
observed value of the EW(Hβ) results from the emission
produced in the present burst superposed on the continuum
generated by the present burst PLUS all the previous star
formation.
4 CONCLUSIONS
We have developed a simple inversion tool that allows to
reconstruct from observed probability density distributions
of some monotonical parameter, its time evolution.
We applied the inversion method to the EW(Hβ) dis-
tribution of a sample of 217 extreme star forming galaxies
from the SCHG. We have shown that, considering the sam-
ple as a whole, its EW(Hβ) evolution is not well described
by a coeval burst model, and that HII Galaxies seem to have
a star formation history that is closer to that predicted by
a continuous star formation model.
The simplest interpretation is that while the observed
emission lines track the present burst, the underlying con-
tinuum contains the whole history of star formation of the
HII galaxy.
Even though HII Galaxies are selected by their strong
emission lines, they are not truly young galaxies. Most of
them have undergone substantial star formation probably
during the previous 100 - 1000 My to the present burst.
The situation changes when the analysis is restricted to
a sub-sample covering a narrow range in metallicities (7.50 <
12 + log O/H < 8.25). In this case the EW(Hβ) evolution
seems well described by a coeval burst model with an IMF
having an upper mass limit around 80 M⊙.
Clearly it would be very interesting to extend the
method to the analysis of larger samples of galaxies like,
e.g., that one provided by the Sloane Digital Sky survey.
5 ACKNOWLEDGMENTS
Sergiy Silich, Daniel Rosa Gonz´alez and Elena Terlevich
gratefully acknowledge financial support from CONACYT,
the Mexican Research Council, through research grants
# 36132-E and # 32186-E. ET, RJT and DRG are grateful
for the hospitality of the IoA, Cambridge, where part of this
work was accomplished.
REFERENCES
Denicol´o, G., Terlevich, R., Terlevich, E.: 2002, MNRAS, 330,
69
Dottori, H.A. : 1981, Ap&SS, 80, 267
Dottori, H.A., Bica, E.L.D. : 1981, A&A, 102, 245
Dufour, R.J., Garnett, D.R., Skillman, E.D., & Shields, G.A.:
1996, in ASP Conf. Ser.,98, From Stars to Galaxies: The Im-
pact of Stellar Physics on Galaxy Evolution, ed. C. Leitherer,
U. Fritze-von Alvensleben, & J. Huchra (San Francisco: ASP),
358
Garnett, D. R.; Skillman, E. D.; Dufour, R. J., & Shields, G. A.:
1997, ApJ, 481, 174
Kunth, D., & Ostlin, G. : 2000 , A& Ap. Rv., 10, 1
Legrand, F., Kunth, D., Roy, J.-R., Mas-Hesse, J.M. and Walsh,
J.R. : 2000, A&A, 355, 891
Leitherer, C. & Heckman, T.M. : 1995 , ApJS, 96, 9
Leitherer, C., Schaerer, D., Goldader, J.D., Gonz´alez-Delgado,
R.M., Robert, C., Kune, D.F., de Mello, D.F., Devost, D., &
Heckman, T.M. : 1999, ApJS, 123, 3
Melnick,J., : 1987,
in Starbursts and Galaxy Evolution, ed.
T.X.Thuan, T.Montmerle and J.Tran Than Van (Edition
Frontieres), 215
Melnick, J. : 1992,
in Star Formation in Stellar systems, eds.
G.Tenorio-Tagle, M.Prieto and F.S´anchez (CUP, Cambridge),
253
Pagel, B.E.J. : 1998, Nucleosynthesis and chemical evolution of
galaxies, C.U.P., Cambridge, U.K.
Pilyugin, L. S. :1999, A&A, 346, 428
Pilyugin, L. S. and Edmunds, M.G. :1996, A&A, 313, 792
Roy, J.-R., and Kunth, D., : 1995, A&A, 294, 432
Scalo, J., and Wheeler, C., : 2001 , ApJ, 562, 66
Silich, S., Tenorio-Tagle, G., Munoz-Tun´on, C., and Cairos, L.
M., : 2002 , Astron. J., 123 , 2438
Stasi´nska, G., and Leitherer : 1996, ApJS, 107, 661
Telles, E., and Terlevich, R. : 1993, AP&SS, 205, 49
Telles, E., & Terlevich, R. : 1995, MNRAS, 275, 1
Telles, E., & Terlevich, R. : 1997, MNRAS, 286, 183
Telles, E., Melnick, J., & Terlevich, R. : 1997 , MNRAS, 288, 78
Terlevich, R. : 1998 ,
in The Post-Recombination universe, ed.
N.Kaiser & A.N.Lasenby (Dordrecht, Kluwer Academic Pub-
lishers), 69
Terlevich, R., Melnick, J., Masegosa, J., Moles, M., & Copetti,
M.V.F. : 1991, A&AS, 91, 285 (SCHG)
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
How old are the HII Galaxies?
7
Figure 5. Metallicity vs. EW(Hβ) from Denicol´o et al. (2002) compilation. The rectangle isolates the selected restricted metallicity
range used for the analysis (see text). It engulfes 70 objects.
Figure 6. The inversion of the time evolution of the EW(Hβ) for the restricted metallicity range subsample. The thick solid line
represents the inversion, and the thin dashed and dotted lines represent model predictions (SB99) with three different upper mass values
and two different slopes for the IMF, as labelled.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
8
Roberto Terlevich et al.
This paper has been produced using the Royal Astronomical
Society/Blackwell Science LATEX style file.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 8
|
astro-ph/0701676 | 1 | 0701 | 2007-01-24T10:51:11 | High-energy Emission from Pulsar Outer Magnetospheres: Two-dimensional Electrodynamics and Phase-averaged Spectra | [
"astro-ph"
] | We investigate particle accelerators in rotating neutron-star magnetospheres, by simultaneously solving the Poisson equation for the electrostatic potential together with the Boltzmann equations for electrons, positrons and photons on the poloidal plane. Applying the scheme to the three pulsars, Crab, Vela and PSR B1951+32, we demonstrate that the observed phase-averaged spectra are basically reproduced from infrared to very high energies. It is found that the Vela's spectrum in 10-50 GeV is sensitive to the three-dimensional magnetic field configuration near the light cylinder; thus, a careful argument is required to discriminate the inner-gap and outer-gap emissions using a gamma-ray telescope like GLAST. It is also found that PSR B1951+32 has a large inverse-Compton flux in TeV energies, which is to be detected by ground-based air Cerenkov telescopes as a pulsed emission. | astro-ph | astro-ph |
High-energy Emission from Pulsar Outer Magnetospheres
Kouichi Hirotani
ASIAA/National Tsing Hua University - TIARA,
PO Box 23-141, Taipei, Taiwan1
[email protected]
ABSTRACT
We investigate particle accelerators in rotating neutron-star magnetospheres, by simultane-
ously solving the Poisson equation for the electrostatic potential together with the Boltzmann
equations for electrons, positrons and photons on the poloidal plane. Applying the scheme to the
three pulsars, Crab, Vela and PSR B1951+32, we demonstrate that the observed phase-averaged
spectra are basically reproduced from infrared to very high energies. It is found that the Vela's
spectrum in 10-50 GeV is sensitive to the three-dimensional magnetic field configuration near the
light cylinder; thus, a careful argument is required to discriminate the inner-gap and outer-gap
emissions using a gamma-ray telescope like GLAST. It is also found that PSR B1951+32 has a
large inverse-Compton flux in TeV energies, which is to be detected by ground-based air Cerenkov
telescopes as a pulsed emission.
theory -- magnetic fields -- methods: numerical -- pulsars:
individ-
Subject headings: gamma-rays:
ual(B1951+32,Crab,Vela)
1.
Introduction
Pulsars form the second most numerous class of
objects detected in high-energy γ-rays. To date,
six have been detected by the Energetic Gamma
Ray Experiment Telescope (EGRET) aboard the
Compton Gamma Ray Observatory. The γ-ray
pulsations observed from these objects are par-
ticularly important as a direct signature of non-
thermal processes in rotating neutron-star magne-
tospheres, and potentially should help to discrim-
inate among different emission models.
The pulsar magnetosphere can be divided into
two zones: The closed zone filled with a dense
plasma corotating with the star, and the open zone
in which plasma flows along the open field lines to
escape through the light cylinder. The last-open
field lines form the border of the open magnetic
field line bundle. In all the pulsar emission mod-
els, particle acceleration takes place in this open
1Postal address: TIARA, Department of Physics, Na-
2, Kuang Fu
tional Tsing Hua University, 101, Sec.
Rd.,Hsinchu, Taiwan 300
1
zone. In inner-gap (IG) models, which adopts par-
ticle acceleration within several neutron-star radii
above the polar-cap surface, the energetics and
pair cascade spectrum have had success in repro-
ducing the observations (e.g., Daugherty & Hard-
ing 1982, 1996). However, the predicted beam
size is too small to produce the wide pulse pro-
files that are observed from high-energy pulsars.
Seeking the possibility of a wide hollow cone emis-
sion due to flaring field lines, Muslimov and Hard-
ing (2004) extended the idea of Arons (1983) and
proposed a slot-gap (SG) model, in which emis-
sion takes place very close to the last-open field
line from the stellar surface to the outer magneto-
sphere. Since the SG model is an extension of the
IG model into the outer magnetosphere, a nega-
tive magnetic-field-aligned electric field, Ek, arises
if the magnetic moment vector points in the same
hemisphere as the rotation vector. However, the
electric current induced by the negative Ek does
not have a self-consistent closure within the model
(Hirotani 2006, hereafter H06).
To contrive an accelerator model that predicts
an outward current in the lower latitudes (within
the open zone),
it is straightforward to extend
outer-gap (OG) models (Cheng et al. 1986; Ro-
mani & Yadigaroglu 1995, hereafter RY95; Cheng
et al. 2000) into the inner magnetosphere. Ex-
tending several OG models (Hirotani et al. 2003;
Takata et al. 2004), H06 demonstrated that the
gap extends from the stellar surface to the outer
magnetosphere,
that the positive Ek extracts
ions from the star as a space-charge-limited flow
(SCLF), and that most photon emission takes
place in the outer magnetosphere because Ek is
highly screened inside the null surface owing to the
discharge of the created pairs. In the present let-
ter, we formulate the scheme (Beskin et al. 1992)
in § 2, apply it to three rotation-powered pulsars
in § 3, and give a brief discussion in § 4.
2. Gap Electrodynamics
We follow the scheme described in § 2 of H06 to
solve the set of Maxwell and Boltzmann equations.
The first equation we have to consider is the Pois-
son equation for the electro-static potential, Ψ. As
space is limited, we present its Newtonian expres-
sion, −∇2Ψ = 4π(ρ − ρGJ). If the real charge den-
sity, ρ, deviates from the Goldreich-Julian (GJ)
charge density, ρGJ, in some region, an acceler-
ation electric field Ek ≡ −∂Ψ/∂s arises, where s
designates the distance along a magnetic field line.
The second equation we have to consider is the
Boltzmann equations for e±'s. Assuming a steady
state in the frame of reference corotating with the
magnetosphere, we obtain
c cos χ
∂n±
∂s
+
dp
dt
∂n±
∂p
+
dχ
dt
∂n±
∂χ
= S±,
(1)
where c denotes the speed of light, n+ (or n−) the
dimensionless positronic (or electronic) distribu-
tion function normalized by the local GJ number
density. The temporal derivatives of momentum
and pitch angle, dp/dt and dχ/dt, and the colli-
sion term S± are explicitly given in H06. Synchro-
curvature radiation-reaction force is included as an
external force in dp/dt, while the effects of inverse-
Compton scatterings (ICS) and (one-photon and
two-photon) pair creation processes are in S±.
The third equation we have to consider is the
Boltzmann equations for photons. Assuming a
steady state, and neglecting azimuthal propaga-
2
tions, we obtain
c
kr
k
∂g
∂r
+ c
kθ
k
∂g
∂θ
= Sγ(r, θ, ck, kr, kθ),
(2)
where the wave numbers kr and kθ are given by
the ray path, (r,θ) are the polar coordinates, and
the dimensionless photon distribution function g is
normalized by the GJ number density at the stel-
lar surface. ICS, synchro-curvature emission, and
the absorption are contained in Sγ. In H06, the
rate of synchrotron emission by secondary pairs
created outside the gap, was calculated assuming
a constant pitch angle. However, it turns out that
only 17 % of the initial particle energy is con-
verted into radiation. In this letter, we corrected
this problem by computing the pitch angle evo-
lution of radiating particles, which increases the
synchrotron cooling time and hence recovers the
time-integrated, radiated energy to 100 % of the
initial particle energy. Note that the gap electro-
dynamics investigated in H06 remains correct, de-
spite the insufficient secondary synchrotron fluxes
in H06.
To solve the Poisson equation, we impose Ψ = 0
at the lower, upper, and inner (s = 0) boundaries,
and Ek = −∂Ψ/∂s = 0 at the outer boundary.
Ion extraction rate is regulated by the condition
Ek = 0 at s = 0. We parameterize the trans-field
thickness of the gap with hm. If hm = 1 the gap
exists along all the open field lines, while if hm ≪ 1
the gap becomes transversely thin.
3. Application to Individual Pulsars
We apply the theory to three γ-ray pulsars,
Crab, Vela, and B1951+32, focusing on their pho-
ton spectra. Even near and outside the light cylin-
der, photon emission and absorption occur effec-
tively; thus, equations (1) and (2) are solved in
0 < s < 16LC, where Ek = 0 in > 0.9LC;
LC ≡ c/Ω denotes the light cylinder radius, Ω
the stellar rotation frequency, and the distance
from the rotation axis. The field line geometry in
0.9LC < < 2LC mimics the aligned dipole
in the force-free magnetosphere (Contopoulos et
al. 1999; Gruzinov 2005). In > 2LC, the field
lines are assumed to be straight so that they con-
nect smoothly at = 2LC.
3.1. Crab pulsar
3.2. Vela pulsar
We present the results for the Crab pulsar in fig-
ure 1, adopting a magnetic inclination of αi = 75◦
and a dipole moment of µ = 4 × 1030 G cm3, which
is close to the value (3.8 × 1030 G cm3) deduced
from the dipole radiation formula. It follows that
the observed pulsed spectrum from IR to VHE
can be reproduced, provided that we observe the
photons emitted into 75◦ < θ < 103◦, where θ de-
notes the photon propagation direction measured
from the rotational axis. Because of the aberra-
tion of light, it is reasonable to suppose that pho-
tons emitted in a certain range of θ comes into
our line of sight in an obliquely rotating three-
dimensional magnetosphere (e.g., RY95). The flux
rapidly decreases with decreasing θ for 75◦ < θ <
93◦, because Ek is highly screened in the inner
part of the gap. Nevertheless, an average over
75◦ < θ < 103◦, which includes negligible flux
between 75◦ < θ < 89◦, achieves the current ob-
jective, because the spectral normalization can be
fitted (within a factor of a few) without changing
the spectral shape, by adjusting hm.
ICS takes place efficiently near and outside the
light cylinder (Aharonian & Bogovalov 2003), be-
cause the magnetospheric IR photons, which are
emitted along convex field lines, collide with the
gap-accelerated positrons near the light cylinder,
where the field lines are concave. Most of such
upscattered photons, as well as some of the high-
energy tail of the curvature component, are ab-
sorbed by the γ-γ collisions. As a result, there is
a gradual turnover around 10 GeV, which forms
a striking contrast with the steep turnover pre-
dicted in IG models due to magnetic pair creation.
The primary curvature component appears be-
tween 100 MeV and 10 GeV, while the secondary
synchrotron component appears below a few MeV.
Between a few MeV and 100 MeV, the ICS compo-
nent dominates, because the secondary pairs that
have been cooled down below a few hundred MeV
efficiently up-scatter magnetospheric UV and X-
rays to lose energy. Similar spectral shapes are
obtained for different values of αi, µ, hm, provided
that the created current is super GJ, by virtue of
the negative feedback effects (H06).
Next, we present the results for the Vela pul-
sar in figure 2 (left). Taking an angle average
over 75◦ < θ < 103◦ (solid line), we can repro-
duce the observed pulsed spectrum, except for the
RXTE results. The primary curvature component
appears between 100 MeV and 10 GeV, while the
secondary and tertiary synchrotron components
appear below 100 MeV. ICS is negligible for the
Vela pulsar because of its weak magnetospheric
emission. Similar spectral shapes are obtained for
super-GJ solutions, even though we have to adjust
hm to obtain an appropriate flux.
In the right panel, we compare the present re-
sults with IG (dotted) and OG (dashed) models,
where the dash-dotted line denotes the averaged
flux for 75◦ < θ < 107◦. It follows that the spec-
trum between 10 and 100 GeV crucially depends
on the angles in which the photons that we ob-
serve are emitted. Thus, to quantitatively predict
the γ-ray emission from the outer magnetosphere,
it is essential to examine the three-dimensional
magnetic field structure near and outside the light
cylinder.
3.3. PSR B1951+32
Thirdly and finally, we present the spectrum
of B1951+32 in figure 3 (left).
It follows that
the ROSAT and EGRET fluxes are reproduced by
taking the flux average in 75◦ < θ < 103◦ (solid
line), except for 17 GeV flux, which was derived
from only two photons. Because of its weak mag-
netic field, less energetic synchrotron photons re-
duce the absorption of the ICS component. The
reduced absorption results in a small synchrotron
flux, which further reduces the absorption outside
the light cylinder, leading to a large, unabsorbed
TeV fluxes. It also follows that the spectrum turns
over at lower energy than the IG model prediction
(dotted curve in the right panel; Harding 2001).
It should be noted that the predicted ICS flux
(above 100 GeV) represents a kind of upper lim-
its, because it is obtained by assuming that all the
magnetospheric synchrotron photons illuminate
the equatorial region in which gap-accelerated
positrons are migrating.
Some of such VHE
photons materialize as energetic secondary pairs,
emitting the synchrotron component between a
few keV and 100 MeV. Between 100 MeV and
3
30 GeV, the primary curvature component domi-
nates, which represents the absolute flux (instead
of upper limits). Some of such curvature photons
have energies above 10 GeV and are efficiently ab-
sorbed to materialize as less energetic pairs, which
emit synchrotron radiation below a few keV. Thus,
the spectrum below a few keV also represents the
absolute flux.
4. Summary and Discussion
To sum up, the self-consistent gap solutions ba-
sically reproduce the observed power-law spectra
below a few GeV for the three pulsars examined.
The cut-off spectra between 10 GeV and 100 GeV
strongly reflect the three-dimensional magnetic
field configuration near the light cylinder; thus, a
discrimination between IG and OG models (e.g.,
using GLAST) should be carefully carried out.
If pulsations are detected above 100 GeV, it un-
doubtedly indicates that the photons are emitted
via ICS near the light cylinder, because VHE emis-
sions cannot be expected in IG models.
Since our analysis is limited within the two-
dimensional plane formed by the magnetic field
lines that thread the stellar surface on the plane
containing both the rotational and magnetic axes,
azimuthal structure is still unknown. Thus, we
cannot present pulse profiles, phase-resolved spec-
tra, or the polarization angle variations in this let-
ter. Since the gap is most active in the outer part
of the magnetosphere (unlike previous OG mod-
els, which adopt the vacuum solution of the Pois-
son equation and hence a uniform emissivity), and
since the photons will be emitted along the in-
stantaneous particle motion measured by a static
observer (unlike the treatment in the OG model
of RY95, who assume a very strong aberration of
light near the light cylinder), it is possible that the
predicted pulse profiles and so on are quite differ-
ent from previous OG models. These topics will
be discussed in the subsequent paper, which deals
with the three-dimensional gap structure near the
light cylinder.
The pulsed TeV flux from PSR B1951+32 can
be predicted to be above 5 × 1010 Jy Hz, pro-
vided that a certain fraction (more than 30%) of
the magnetospheric soft photons illuminate the
equatorial region. However, if the poloidal field
4
lines are more or less straight near the light cylin-
der, as demonstrated by Spitkovsky (2006, fig. 2)
for an oblique rotator, the equatorial region may
not be efficiently illuminated.
In this case, the
VHE flux will decrease from the current predic-
tion. There is room for further investigation how
to extend the present analysis into three dimen-
sions, and to combine it with time-dependence
three-dimensional force-free electrodynamics.
The author is grateful to Drs.
N. Otte,
R. Taam, J. Takata for helpful suggestions. This
work is supported by the Theoretical Institute for
Advanced Research in Astrophysics (TIARA) op-
erated under Academia Sinica and the National
Science Council Excellence Projects program in
Taiwan administered through grant number NSC
95-2752-M-007-006-PAE.
REFERENCES
Aharonian, F. A., Bogovalov, S. V. 2003, New As-
tronomy 8, 85
Arons, J. 1983, ApJ 302, 301
Becker, W., & Trumper, J. 1996, A&AS 120, 69
Beskin, V. S., Istomin, Ya. N., & Par'ev, V. I.
1992, Sov. Astron. 36(6), 642
Borione, A. et al. 1997, ApJ 481, 313
Chang, H.-K., Ho, C. 1997 ApJ479, L125
Cheng, K. S., Ho, C., & Ruderman, M. 1986, ApJ,
300, 500 (CHR86)
Cheng, K. S., & Zhang, L. 1996, ApJ463, 271
Cheng, K. S., Ruderman, M., & Zhang, L. 2000,
ApJ, 537, 964
Clifton, T. R., Backer, D. C., Neugebauer, G. et
al. 1988, A&Ap 191, L7
Contopoulos, I., Kazanas, D., and Fendt, C. 1999,
ApJ, 511, 351
Daugherty, J. K., & Harding, A. K. 1982, ApJ,
252, 337
Daugherty, J. K., & Harding, A. K. 1996, ApJ,
458, 278
de Naurois, M. et al. 2002, ApJ566, 343
Ramanamurthy, P. V. et al. 1995b, ApJ450, 791
Eikenberry, S. S., Fazio, G. G., Ransom, S. M. et
Romani, R. W. 1996, ApJ, 470, 469
al. 1997, ApJ 477, 465
Romani, R. W., & Yadigaroglu,
I. A. 1995,
Fierro, J. M., Michelson, P. F., Nolan, P. L., &
ApJ438, 314
Thompson, D. J., 1998, ApJ494, 734
Safi-Harb, S. Ogelman, H., Finley, J. P. 1995,
Goldreich, P. Julian, W. H. 1969, ApJ. 157, 869
ApJ439, 722
Gruzinov, A. 2005, Phys. Rev. Letters 94, 021101
Spitkovsky, A. 2006, ApJL 648, 51
Srinivasan, R. et al. 1997, ApJ489, 170
Takata, J., Shibata, S., & Hirotani, K., 2004, MN-
RAS354, 1120
Tanimori et al. 1998, ApJ 492, L33
Thompson, D. J. et al. 1999, ApJ516, 297
Ulmer, M. P. et al. 1995, ApJ448, 356
Weisskopf, M., O'Dell, S. L., Paerels, F. et al.
2004, ApJ 601, 1057
Harding, A. K. & Daugherty, J. K. 1993, in proc.
of the Los Alamos workshop on isolated pulsars,
ed. van Riper, K. A. et al., Cambridge Univ.
Press, p. 279
Harding, A. K. 2001, American Institute of
Physics (AIP) Proceedings, 558, 115
Harding, A. K., Strickman, M. S., Gwinn, C. et
al. 2002, ApJ, 576, 376
Hillas, A. M., Akerlof, C. W., Biller, S. D. et al.
1998, ApJ503, 744
Hirotani, K., Harding, A. K., & Shibata, S., 2003,
ApJ591, 334
Hirotani, K. 2006, ApJ652, 1475 (H06)
Kanbach, G. et al. 1994, A & A 289, 855
Knight, F. K. 1982, ApJ 260, 538
Kuiper, L., Hermsen, W., Cusumano, G. et al.
2001, A& A 378, 918
Lessard, R. W. et al. 2000, ApJ 531, 942
Manchester, R. N., Wallace, P. T., Peterson, B.
A., Elliott, K. H. 1980, 190, 9
Mestel, L. 1971, Nature Phys. Sci., 233, 149
Mineo, T., Ferrigno, C., Foschini, L. et al. 2006,
A&A, 450, 617
Much, R. et al. 1995, A&A 299, 435
Muslimov, A. G., & Harding, A. K., 2004, ApJ,
606, 1143 (MH04)
Nolan, P. L. et al. 1993, ApJ409, 697
Percival, J. W. et al. 1993, ApJ 407, 276
Ramanamurthy, P. V. et al. 1995a, ApJ447, L109
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
5
Crab
bestfit (q <103o)
Fig. 1. -- Phase-averaged spectrum of the mag-
netospheric emissions from the Crab pulsar for
αi = 75◦, µ = 4 × 1030 G cm3 and hm = 0.12. The
thick dashed, dash-dotted, dotted and dash -- triple-
dotted lines represent fluxes into 93◦ < θ < 97◦,
97◦ < θ < 101◦, 101◦ < θ < 105◦,105◦ < θ <
109◦; the thin dashed, dash-dotted, dotted ones
109◦ < θ < 113◦,113◦ < θ < 117◦, 117◦ < θ <
121◦. The thick solid line represents the averaged
flux in 75◦ < θ < 103◦. See Eikenberry et al.
(1997) for IR-UV data; Knight (1982), Weisskopf
et al. (2004), and Mineo et al. (2006) for X-rays;
Nolan et al. (1993), Ulmer et al. (1995), Much et
al. (1995), Fierro et al. (1998), Kuiper et al. (2001)
for 10 MeV -- 20GeV; Borione et al. (1997), Tan-
imori et al. (1998), Hillas et al. (1998), Lessard
et al. (2000), and de Naurois et al. (2002) for the
upper limits above 50 GeV.
6
Vela
bestfit (q
< 103o)
Vela
< 103o)
< 107o)
this work (q
this work (q
OG model
IG model
Fig. 2. -- Phase-averaged Vela spectrum for µ =
4 × 1030 G cm3, αi = 75◦ and hm = 0.21. See
Manchester et al. (1980) for optical data (open
squares), Harding et al. (2002) for X-rays (filled
triangles), Kanbach et al. (1994), Ramanamurthy
et al. (1995b), Fierro et al. (1998), Thompson et
al. (1999) for γ-rays (open circles). Left: Same
figure as figure 1. Right: Close up above 100 MeV
to compare with IG (Harding & Daugherty 1993)
and traditional OG (Romani 1996) models. The
solid line is same as the left panel.
7
bestfit (q
< 103o)
EGRET
+ 75GeV cutoff
this work
(q
< 103o)
this work
(q
< 107o)
IG model
3. --
Fig.
Phase-averaged spectrum of
PSR B1951+32 for µ = 2 × 1029 G cm3, αi = 75◦
and hm = 0.39. Lines represent same quanti-
ties as figure 2 unless notified. See Clifton et
al.
(1988) for IR upper limits (open squares),
Becker & Trumper (1996), Safi-Harb et al. (1995)
and Chang & Ho (1997) for X-ray data (open
triangles), Ramanamurthy et al.
for
100 MeV-15 GeV data (open circles), Srinivasan
et al. (1997) for VHE upper limits.
(1995a),
8
|
astro-ph/0604445 | 1 | 0604 | 2006-04-20T20:27:24 | Low-Mass Binary Induced Outflows from Asymptotic Giant Branch Stars | [
"astro-ph"
] | A significant fraction of planetary nebulae (PNe) and proto-planetary nebulae (PPNe) exhibit aspherical, axisymmetric structures, many of which are highly collimated. The origin of these structures is not entirely understood, however recent evidence suggests that many observed PNe harbor binary systems, which may play a role in their shaping. In an effort to understand how binaries may produce such asymmetries, we study the effect of low-mass (< 0.3 M_sun) companions (planets, brown dwarfs and low-mass main sequence stars) embedded into the envelope of a 3.0 M_sun star during three epochs of its evolution (Red Giant Branch, Asymptotic Giant Branch (AGB), interpulse AGB). We find that common envelope evolution can lead to three qualitatively different consequences: (i) direct ejection of envelope material resulting in a predominately equatorial outflow, (ii) spin-up of the envelope resulting in the possibility of powering an explosive dynamo driven jet and (iii) tidal shredding of the companion into a disc which facilitates a disc driven jet. We study how these features depend on the secondary's mass and discuss observational consequences. | astro-ph | astro-ph | Low-Mass Binary Induced Outflows from Asymptotic Giant
Branch Stars
J. Nordhaus1,2, E. G. Blackman1,2
1. Dept. of Physics and Astronomy, Univ. of Rochester, Rochester, NY 14627
2. Laboratory for Laser Energetics, Univ. of Rochester, Rochester, NY 14623
ABSTRACT
A significant fraction of planetary nebulae (PNe) and proto-planetary
nebulae (PPNe) exhibit aspherical, axisymmetric structures, many of which are
highly collimated. The origin of these structures is not entirely understood,
however recent evidence suggests that many observed PNe harbor binary
systems, which may play a role in their shaping. In an effort to understand
how binaries may produce such asymmetries, we study the effect of low-mass
(< 0.3 M⊙) companions (planets, brown dwarfs and low-mass main sequence
stars) embedded into the envelope of a 3.0 M⊙ star during three epochs of its
evolution (Red Giant Branch, Asymptotic Giant Branch (AGB), interpulse
AGB). We find that common envelope evolution can lead to three qualitatively
different consequences: (i) direct ejection of envelope material resulting in a
predominately equatorial outflow, (ii) spin-up of the envelope resulting in the
possibility of powering an explosive dynamo driven jet and (iii) tidal shredding
of the companion into a disc which facilitates a disc driven jet. We study
how these features depend on the secondary's mass and discuss observational
consequences.
Subject headings: planetary nebulae: general -- stars: AGB and post-AGB --
stars: low-mass, brown dwarfs -- stars: magnetic fields
6
0
0
2
r
p
A
0
2
1
v
5
4
4
4
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
Deviation from spherical symmetry in planetary nebulae (PNe) and protoplanetary
nebular (PPNe) can be pronounced (Soker (1997), Balick & Frank 2002, Bujarrabal et al.
2004). Understanding the origin of the asymmetries is an ongoing aim of current research.
A variety of scenarios have been proposed to explain the transition from progenitor
to planetary nebula. As low- and intermediate-mass main sequence stars evolve onto the
-- 2 --
Asymptotic Giant Branch (AGB), enhanced stellar wind mass loss leads to depletion of the
hydrogen envelope around the central core. Recent surveys (Sahai (2000), Sahai (2002)) of
AGB and post-AGB stars have revealed spherical symmetry leading to the conclusion that
any shaping process must occur in a relatively short time just before the birth of the PNe
in the PPNe, or AGB phase (Bujarrabal et al. 2001).
Bipolar outflows in either the AGB or post-AGB phase could produce such structures.
However, the mechanism by which the bipolar winds are produced remains to be fully
understood. Binary interactions, large-scale magnetic fields and high rotation rates of
isolated AGB stars may all play some role in explaining the observed structures: Frank et al.
(1994) appealed to a superwind induced by a thermal helium flash as a possible production
mechanism for bipolar planetary nebulae. Soker (2002) argued that such a model does not
account for the observational link between aspherical mass loss and asymptotic wind. Such
an observational correlation could be explained by a binary system in which the companion
is a low-mass star or brown dwarf (Soker (2004)).
AGB and post-AGB remnant central stars are known to possess magnetic fields (Bains
et al. (2004) and Jordan et al. (2005)). In addition, direct evidence of a magnetically
collimated jet in an evolved AGB star has been detected (Vlemmings et al. (2006)), further
suggesting some dynamical role for magnetic fields. Magnetic outflows from single stars have
been proposed as mechanisms for shaping PPNe and PNe (Pascoli 1993; Blackman et al.
2001a). However, single star models may be unable to sustain the necessary Poynting flux
required to maintain an outflow through the lifetime of the AGB phase unless differential
rotation is reseeded by convection or supplied by a binary (Blackman 2004, Soker (2006a)).
A model in which a disc driven magnetic dynamo driven outflow is sustained by accretion
from a shredded secondary was pursued in (Blackman et al. 2001b).
In this respect, it is noteworthy that recent studies support the claim that most, if
not all, observed planetary nebulae are the result of a binary interaction (De Marco et al.
(2004), Sorensen & Pollacco (2004), De Marco & Moe (2005), Mauron & Huggins (2006)).
While this conclusion is based on observations and population synthesis studies, Soker
(2006b) points out that the corresponding formation rate of PNe from such studies is 1/3
the formation rate of white dwarfs. This supports the claim that binary stars may produce
the more prominently observed planetary nebulae (Soker & Subag (2005)).
The question of just how a binary shapes a PNe or PPNe remains a topic of active
research. In this paper, we explore the effects of an embedded low-mass companion inside
the envelope of a 3 M⊙ star during three epochs of its evolution off the main sequence (Red
Giant Branch (RGB), AGB and interpulse AGB). A common envelope (CE) facilitates
mass ejection in several ways: The in-spiral of the secondary toward the core deposits
-- 3 --
orbital energy and angular momentum in the envelope. This directly ejects and/or spins up
the envelope. In the latter case, any enhanced differential rotation could aid in magnetic
field generation, which in turn could drive mass loss. In section 2, we describe the stellar
models used and derive the basic equations for in-spiral and for the transfer of energy and
angular momenutum from the secondary to the envelope. Results for different evolutionary
epochs are discussed in section 3. We present observational implications and applications
to specific systems in section 4 and conclude in section 5.
2. Common Envelope Evolution
Under certain conditions, Roche lobe overflow in close binary systems results in both
companions immersed in a CE (Paczynski (1976), Iben & Livio (1993)). Once inside,
velocity differences between companion and envelope generate a drag force that acts to
reduce the orbital separation of the companion and core. Orbital energy is deposited into
the envelope during the in-spiral process. Some of this energy is radiated away while the
rest is available to reduce the gravitational binding energy of the envelope. The efficiency
with which orbital energy unbinds envelope matter is of central importance to CE evolution.
This is commonly incorporated into a parameter, α, which represents the fraction of orbital
energy available for mass ejection as follows:
Ebind = α∆Eorb,
(1)
where ∆Eorb is the change in orbital energy of the binary and Ebind is the energy required to
unbind envelope material. In principle, knowledge of the binding energy and α determines
how much material is ejected, and in the case of complete envelope ejection, final binary
separation distances. Such studies have been performed for a variety of different systems
under a range of conditions of which the following are a small sample (Yungelson et al.
(1995), Dewi & Tauris (2000), Taam & Sandquist(2000), Politano (2004)).
We investigate the effect of embedding planets, brown dwarfs and low-mass main
sequence stars into an envelope of a 3 M⊙ star during various epochs of its evolution off of
the main sequence. That the secondary mass represents a small perturbation to the initial
envelope configuration allows us to neglect detailed radiative and hydrodynamical effects.
We present as simplified a picture as possible in order to elucidate basic phenomonological
consequences of the interaction.
-- 4 --
2.1. Envelope Binding Energy
Our stellar model consists of a 3 M⊙ main sequence star whose evolution is followed
through the AGB phase with X = 0.74 (mass fraction of hydrogen), Y = 0.24 (mass
fraction of helium), Z = 0.02 and no mass loss (S. Kawaler - personal communication, Fig.
1). A range of evolutionary models allows consideration of various positions and times at
which the expanding envelope engulfs the orbiting brown dwarf. We focus on three main
epochs in the evolution: (i) near the tip of the first Red Giant Branch, (ii) the beginning
of the Asymptotic Giant Branch and (iii) the quiet period between thermal pulses on the
AGB branch. In each case, we calculate the energy required to unbind the envelope mass
above a given radius, r (measured from the center of the primary's core) as follows:
Ebind(r) = −Z MT
M
GM(r)
r
dm(r),
(2)
where MT is the total mass of the star and M is the mass interior to the companions
orbital radius. Here it is assumed that the core and envelope do not exchange energy
during the CE phase and that ejection of material has no bearing on core structure. The
values we determine for the binding energy in all three epochs are comparable to results
from an estimation method first proposed by Webbink (1984) and further refined by Dewi
& Tauris (2000) and Tauris & Dewi (2001). Explicitly calculating the binding energy for
each evolutionary epoch fixes our efficiency parameter α between 0 and 1.
For the RGB star, our model core radius rc ∼ 3.5 × 109 cm with the envelope extending
out to a radius r⋆ ∼ 7 × 1011 cm. At the chosen time in the RGB phase, the core contains
0.41 M⊙ and the energy required to unbind the entire envelope (∼ 1048 ergs) is the largest of
our three epochs. Once the star has ascended onto the AGB, the core contracts to a radius
of 2.9 × 109 cm and the envelope expands to r⋆ = 5.7 × 1012 cm. The core has increased
its mass to 0.55 M⊙ and the energy required to unbind the entire envelope decreases to 1.3
× 1047 ergs. For the interpulse AGB phase, the core has contracted to a rc ∼ 1.6 × 109 cm
while the envelope has expanded to r∗ ∼ 1.3 × 1013 cm. The envelope binding energy has
been further reduced to 5.7 × 1046 ergs with the core containing 0.58 M⊙. We find that
the interpulse AGB phase is most favorable for binary induced envelope ejection since the
range of masses and radii required to deposit favorable orbital energy into the envelope is
greatest in this phase. We discuss these results in detail in section 3.
2.2. Orbital Energy and Angular Momentum Evolution
The immersion of the secondary in the envelope of the giant results in a reduction
of the separation distance between core and companion. To calculate the in-spiral and
-- 5 --
angular momentum transfer, we need to equate the rate of energy lost by drag to the
change in gravitational potential energy. The motion of a body under the influence of a
central potential while incurring a drag force has been well studied and a general set of
equations can be found in Pollard (1976). Here we limit ourselves to the case where orbital
eccentricity is negligible, such that the planet exhibits approximate Keplerian motion at
each radii. Under these conditions, the energy per unit time released by the secondary mass
takes the following form:
Ldrag = ξπR2
aρ(v − venv)3
,
(3)
where v = (vr, vφ, 0) is the companion velocity, ρ the envelope density, venv the envelope
velocity, Ra the accretion radius measured from the center of the companion, and ξ is a
dimensionless factor dependent upon the Mach number of the companions motion with
respect to the envelope. For supersonic motion, ξ is greater than or equal to 2 (Shima et
al. (1985)). The orbital motion of the planet is supersonic everywhere inside the orbit (see
Fig. 1) and for simplicity, we assume a value of ξ = 4. The value of ξ acts only to slightly
increase or decrease the in-spiral time of the secondary, while leaving the underlying physics
unchanged. The accretion radius is then given by
Ra ∼
2Gm2
(v − venv)2 ,
(4)
and represents the region around the secondary inside of which matter is gravitationally
attracted to the secondary as it passes through the envelope. If the companion moves close
enough to the core, tidal effects can shred it. We estimate the shredding radius (measured
from the center of the primay's core) from balancing the differential gravitational force
across the size of the companion R2 (measured from the center of the companion) with its
self gravity, that is, d
R2.
dr (cid:16) GM
r2 (cid:17) R2 ≃ Gm2
R2
, which yields rs ≃ 3q 2M
m2
2
To determine the companion size, R2, we separate our objects into three distinct
groups: planets (m2 ≤ 0.0026 M⊙; Zapolski & Salpeter (1969)), brown dwarfs (0.0026
M⊙ < m2 < 0.077 M⊙; Burrows et al. (1993)) and low-mass main sequence stars. We
adopt an approximation to the models of Burrows et al. (1993), identical to that used in
Reyes-Ruiz & Lopez (1999) for brown dwarfs, namely
R2 = (cid:20)0.117 − 0.054Log2(cid:18) m2
0.0026(cid:19) + 0.024Log3(cid:18) m2
0.0026(cid:19)(cid:21) R⊙.
(5)
For low-mass main sequence stars, we adopt the homologous power-law used in Reyes-Ruiz
& Lopez (1999) given by
R2 = m2
M⊙!0.92
R⊙.
(6)
-- 6 --
As the separation between core and companion decreases, the secondary begins to fill its
Roche lobe. An approximation for the Roche lobe radius is given by (Eggleton (1983))
RRL =
0.49q2/3r
0.6q2/3 + Ln(1 + q1/3)
,
(7)
where q is the ratio of secondary mass to core mass and r is the binary separation distance.
Once RRL = R2, mass transfer ensues.
The time rate of change of gravitational potential energy of the binary is given by:
dU
dt
=
Gm2vr
r
dM
dr
−
M
r ! .
(8)
As the secondary traverses the envelope, the drag luminosity must be supplied by the
change in gravitational potential energy. Therefore, we equate (3) and (8) using (4) and
obtain an equation for the infall velocity. This yields
vr =
4ξπGm2rρ
,
(9)
(cid:16) dM
dr − M
r (cid:17)qv2
r + ¯v2
φ
where ¯vφ = vφ − venv ≃ vφ for slowly rotating stars. In addition, vr ≪ vφ everywhere inside
the envelope. Eq. (9) agrees with the limit of a general set of equations found in Alexander
et al. (1976) under these circumstances. The time scale for infall from a position inside
the envelope can then be estimated as τ ∼ r
(see Fig. 2). The in-spiral time is slightly
vr
shorter for the AGB star than for the interpulse AGB star. In the outer reaches of the
envelope, τ is comparable to the lifetime on the AGB (∼ 105 yrs), but sharply drops to ∼ 1
yr just inside the outer radius.
As a consequence of the in-spiral process, the secondary loses orbital energy and
angular momentum. The reduction in orbital energy is given by:
∆Eorb(r) =
GMT m2
2ro
−
GM m2
2r
,
(10)
where ro ∼ r⋆ is the stellar radius. In practice, we expect ro to be slightly less then r⋆ since
material in the outer reaches of the envelope exerts little drag force, thereby significantly
extending the infall time (see Fig. 2). In addition to transfer of orbital energy, as the
secondary moves closer to the core, conservation of angular momentum results in a spin up
of the initially stationary envelope. We assume that the lost orbital angular momentum is
transferred to spherical shells of the envelope. In reality, most of the angular momentum
may be confined close to the orbital plane resulting in latitudinal differential rotation
in addition to that expected in the radial direction. A more equatorially concentrated
-- 7 --
deposition of angular momentum could therefore lead to even more differential rotation
than considered herein, and a more robust dynamo.
The simple equations in this section allow us to crudely investigate different outcomes
of CE evolution. Equations (1), (2), (5), (6) and (10) determine the depth at which various
mass secondaries can penetrate into the star before depositing enough energy to unbind
envelope material and Eq. (9) gives the radial component of velocity during in-spiral. In the
next section, we discuss different CE end states that result from analyzing these equations.
3. Common Envelope Evolution Scenarios
We outline three qualitatively different scenarios that can occur once the secondary
is immersed in the envelope of a given stellar evolution phase: (i) the secondary provides
enough orbital energy to directly unbind the envelope (or a portion of it) by itself, (ii) the
secondary induces differential rotation in the envelope which can power a dynamo therein,
unbinding the envelope or (iii) the secondary is shredded into a disc around the core, which
can lead to a disc driven outflow. These scenarios are presented schematically in Fig. 3.
Below we discuss each in depth and comment on their observational implications in Sec. 4.
3.1. Secondary Induced Envelope Expulsion
As the secondary enters the primary's envelope, the mutual drag transfers angular
momentum to the latter and the secondary spirals in. For an RGB star, envelope accretion
onto brown dwarf secondaries was previously studied (Livio & Soker (1984), Soker et al.
(1984)). The stellar model consisted of a 0.88 M⊙ giant with a core mass of 0.72 M⊙ during
hydrogen and helium shell burning phase. The evolution of the giant was subsequently
followed during in-spiral. The authors found that the secondary evaporated if its mass was
below an initial critical value (mcrit ≃ 0.005 M⊙). When the secondary exceeded this mass,
it instead grew to 0.14 M⊙, independent of its initial supercritical mass. The evolution was
followed until the envelope was ejected, leaving a close binary system.
For each evolutionary phase and fixed value of α, we calculate the radius at which the
orbital energy released equals that of the binding energy of the envelope for a range of
secondary masses (see Fig. 4).
When the star has just reached the RGB phase, we find that no objects under 0.5 M⊙
can unbind the envelope before they are tidally shredded. Even if the companion is not
shredded, extracting enough orbital energy to expel the envelope requires penetrating to
-- 8 --
the inner most regions where physical contact with the core results. Thus, we do not expect
low-mass secondaries inside our model RGB star to produce helium white dwarfs.
As the star enters its AGB phase, the envelope expands and the core contracts, thereby
lowering the binding energy. In this case, we find a range of masses and efficiencies for
which the CE interaction can expel the envelope directly (without magnetic fields). A 0.15
M⊙ companion provides enough orbital energy (for α = 0.4) to unbind the envelope when it
reaches a radius of 3 × 1010 cm. For the interpulse AGB star, the binding energy is further
reduced from the initial AGB, extending the range of masses and drag efficiencies that can
unbind the system. Even if only 20 percent of the orbital energy is available for envelope
ejection (α = 0.2), a 0.2 M⊙ secondary can expel the envelope at 5 × 1010 cm. The binding
energy as a function of position is shown in Fig. 5. In addition, the orbital energy that a
0.02 M⊙ companion supplies as it traverses the envelope is also shown.
3.2. Secondary Induced Envelope α − Ω Dynamo
Outflow production mechanisms that extract rotational energy may be required to
explain the observed high power bipolar features of PN and PPNe, since radiation driving
is insufficient (Bujarrabal 2001). Magnetic field generation inside the AGB envelope may
provide a way of extracting this energy and collimating an outflow. Magnetically mediated
outflows in an isolated AGB star have been studied from different perspectives (e.g. Pascoli
1993, Blackman et al. (2001a), Soker & Zoabi (2002)) and outflows from binary + disc
systems (Reyes-Ruiz & Lopez (1999), Regos & Tout (1995), Soker & Livio (1994), Blackman
et al. 2001b) have also been considered. Here we focus on outflows from binary induced
dynamos in the stellar envelope, and discuss accretion driven outflows in the next section.
The AGB phase of stellar evolution provides the conditions needed to power an
α − Ω dynamo analagous to those studied in the sun, white dwarfs, and supernova
progenitors (Parker (1993), Thomas et al. (1995), Blackman, Nordhaus & Thomas (2006)).
The combination of a deep convective envelope and differential rotation could generate
large-scale magnetic fields. Blackman et al. (2001a) investigated an interface dynamo in
the context of our 3.0 M⊙ AGB model. They assumed that the star was initially rotating
on the main sequence with a rotation rate of 200 km/s. Assuming that evolution off the
main sequence conserves angular momentum on spherical shells results in a differentially
rotating AGB star (see Fig. 6). Large-scale saturated fields of B ∼ 5 × 104 G can then be
calculated at the base of the convection zone. But to drive a magnetic outflow, the dynamo
must operate over the entire lifetime of the AGB phase (∼ 105 yrs) until enough matter has
been radiatively bled from the envelope for the magnetic "spring" like a jack-in-the-box.
-- 9 --
Unfortunately, field amplification drains energy from differential rotation and acts to
transfer angular momentum from the core to the envelope, slowing down the core within 100
yrs. Anisotropic convection does provide a possible mechanism for reseeding the differential
rotation (c.f. Rudiger 1989) and maintaining a steady AGB rotation profile (Blackman
2004), but more work is needed to assess the viability of the single star mechanism.
Alternatively, a binary companion may, via a CE phase, supply enough differential
rotation to the envelope that the resulting amplified Poynting flux is large enough to
unbind the envelope within a few years (Blackman 2004). We study this concept more
carefully here. During the CE, the transfer of angular momentum and orbital energy to
the envelope induces in-spiral of the companion. Fig. 3 shows the envelope rotation profile
produced when a 0.02 M⊙ brown dwarf travels from the outer part of the envelope to the
core boundary at the beginning of the AGB phase. The differential rotation profile from
the single star approach of Blackman et al. (2001a) is presented for comparison. The
magnitude of rotation generated from a binary interaction during the CE phase is a factor
of 10 greater in the interface region and can therefore supply a significantly larger amount
of differential rotation energy for a dynamo. In principle, if the secondary could penetrate
all the way to the core-envelope boundary (Fig. 5), an additional region of energy could
be tapped. However, the penetration depth of the secondary is limited by tidal shredding,
while the energy available for field amplification is constrained by how far the poloidal field
can diffuse into the differential rotation zone (see Blackman, Nordhaus & Thomas (2006)
for details).
It should be noted that an interface dynamo will rapidly drain the available differential
rotation energy (Blackman, Nordhaus, Thomas 2006) so unless the differential rotation is
reseeded by convection, the outflow produced by such a dynamo would be explosive and
last < 100 years. This is suggestive of ansae further comment on this in section 4.2.
For lower mass companions, the transfer of angular momentum may not produce
strong enough differential rotation to drive a robust dynamo. The effect of a modest AGB
dynamo (Soker (2001b)) might be to produce more sunspots near the stellar equator. Dust
formation is increased near these cool spots, enhancing the radiative mass loss rate there.
If such a modest dynamo could last long enough, the formation of elliptical PNe might be
aided by this mechanism.
Further study of an α − Ω interface dynamo produced from a secondary induced
rotation profile is warranted (and in progress). Preliminary results from our investigation
of an interface dynamo operating in the AGB star model are encouraging. For a differential
rotation profile generated by the in-spiral of a 0.02 M⊙ companion (Fig. 6), we obtain cycle
periods of ∼ 0.1 yr with the transient dynamo lasting 0.5 − 1 yr.
-- 10 --
3.3. Disc Driven Outflow
In the event that the secondary cannot supply enough orbital energy to directly unbind
the envelope or spin it up enough to power a dynamo, the secondary will be shredded from
tidal forces as it nears the core. The companion's physical radius, R2, expands to fill its
Roche lobe, at which point, mass transfer to the envelope ensues. Eventually, near the core,
tidal shredding occurs. After several dynamical periods, the remnant secondary mass forms
an accretion disc which may be capable of producing collimated outflows (Blackman et al.
(2001b)). This scenario differs from Morris (1987) in which the secondary strips material
from the AGB primary and acquires a disc.
Soker & Livio (1994) investigated disc formation scenarios and found that a disc could
form around the primary core when a ∼ 1M⊙ main sequence secondary is embedded in an
AGB envelope. At the end of the CE phase, after the primary has shed its envelope, the
secondary expands, loses matter and forms a disc around the primary. This is qualitatively
similar to disc formation in cataclysmic variables, in which matter is stripped off a low-mass
companion and forms a disc around a compact primary. Such a disc may be able to power
collimated outflows during the proto-PNe phase. This situation can occur if the companion
directly ejected much of the envelope and avoided tidal shredding. In the present work,
we focus only on low-mass companions, and on discs formed inside the CE from tidally
destruction of this companion.
Reyes-Ruiz & Lopez (1999) investigated initial binary configurations which lead to disc
formation inside an AGB envelope from Roche lobe overflow of companions with masses
between 0.001 − 1.0 M⊙. Matter flowing through the inner Lagrangian point falls inward
unless it has enough angular momentum to remain in Keplerian orbit. For secondary masses
above 0.05 M⊙, matter stripped off the secondary falls all the way to the core surface and
therefore does not form an accretion disc. For massive planets and smaller brown dwarfs,
an accretion disc can form.
Reyes-Ruiz & Lopez (1999) also study, the evolution of the resulting disc. They find
that a geometrically thin accretion disc forms after an initial mass redistribution stage
of ∼ 1yr. For a 0.03 M⊙ brown dwarf secondary with ≥ 10% of its mass forming a disc,
secondary, the resulting disc is thinner and cooler with the accretion rate dropping to
the accretion rate is found to be ∼ 3 × 10−7h(cid:16) t
4 × 10−8h(cid:16) t
4i M⊙ yr−1. For a 3 Jupiter mass
4i M⊙ yr−1. These estimates for mass flow are comparable to those of
young stellar objects where an accretion disc outflow connection has been established for
similar values (Hartigan et al. (1995)).
5
104(cid:17) −
5
104(cid:17) −
In Fig. 2, we show the distance from the core at which various mass secondaries fill
-- 11 --
their Roche lobes or shred for our AGB and interpulse AGB model stars. If the orbital
energy deposited in the envelope is insufficient to unbid it, then the secondary continues
migrating toward the core. Tidal effects become increasingly important and we expect
objects that penetrate deep enough to be tidally shredded into a disc (see Fig. 4). Brown
dwarfs and massive planets shred to form a disc while low-mass stellar companions, because
of their large radii (∼ 2 × 1010 cm), may contact the core directly before forming a disc.
If the shredding of the companion results in an accretion disc (see Fig. 3), a disk
outflow similar to those observed in other astrophysical objects such as young stellar
objects, X-ray binaries and active galactic nuclei is possible.
Disc driven outflows can sustain their observable lifetime longer than the interface
dynamos. Thus extended bipolarities extending from the PPNe phase into the PNe phase
are suggested of disc mediated outflows rather than merely the explosive outflow of an
interface type dynamo discussed above.
4. Discussion of Observational Implications
As a consequence of our model, outflow composition, collimation and direction vary
based on the mass of the secondary embedded in the envelope. In section 4.1, we discuss
observational implications of our three ejection scenarios (Fig. 3) and the possibility that
they could operate in conjunction. In section 4.2, we comment on specific PPNe and PNe
in the context of our CE framework.
4.1. Observational Consequences
The three basic ejection scenarios are shown in Fig. 3. When the envelope is ejected
purely via orbital energy deposition from the secondary, the corresponding PPNe outflows
will reside primarily in the equatorial plane (see Fig. 3a). There is numerical evidence from
simulations that a binary induced equatorial outflow is confined to an opening angle of ∼ 20
- 30 degrees (Terman & Taam (1996)). Sandquist et al. (1998) follow the three-dimensional
hydrodynamical evolution of an AGB star with companions of 0.4 − 0.6 M⊙. The binary
interaction funnels material and expels it along the equator. If the rotation axis of the
central star can be determined, then the identification of an equatorial outflow suggests a
CE origin.
Many PNe exhibit equatorial tori (Su et al.
(2003), Bujarrabal et al. (1998),
Castro-Carrizo et al. (2002)), some of which are falling back towards the core. This could
-- 12 --
be explained by a CE interaction that did not supply enough energy to unbind the envelope.
Note that very small companions fall into the core without much envelope ejection. For very
large mass companions, the radius at which the envelope is expelled increases, also resulting
in a small amount of matter in the equatorial outflow. There is therefore an intermediate
value of the companion mass which maximizes the amount of equatorial ejecta.
As compared to a high-mass, high density torus, a low-mass, low density torus may
provide inadequate shielding from the central, illuminated white dwarf. As a consequence,
molecule formation in the equatorial torus is reduced. A companion that expels the entire
envelope would create more shielding of the outer parts of the torus, leading to more
molecule survival.
In the event that the secondary cannot directly unbind material, the companion may
induce differential rotation which ejects the envelope via a magnetic dynamo driven outflow
(Fig. 3b). Such an outflow would be predominantly poloidal, likely collimated and aligned
with the central rotation axis. The launching and shaping of the outflow could occur close
to the central core and the role of a magnetic field at larger distances may be less important.
In addition, a torioidal magnetic pressure "sandwich" across the equator acould squeeze
some material out equatorially (Matt et al. 2004). The overall outflow expected from a
magnetic outflow is thus predominantly poloidal with a smaller equatorial component.
If the secondary is shredded into an accretion disc around the core, a disc driven
outflow is possible (Fig. 3c). In this case, the outflow may exhibit chemical signatures of
the destroyed secondary. The atmospheres of brown dwarfs are oxygen-rich, in contrast
to certain carbon-rich AGB stars. Water and carbon monoxide are present in a range of
brown dwarf classes (Geballe et al. (2002), Burgasser et al. (2002)). If the companion is a
brown dwarf, a disc driven outflow can expel oxygen-rich material along the poles. This
may lead to the formation of crystalline silicates along the rotational axis. If the secondary
is a low-mass main sequence star, it may be difficult to detect any difference in outflow
material if the primary is of similar composition.
As the CE phase evolves, a combination of the three above scenarios might occur. For
instance, differential rotation supplied during the CE phase may trigger a dynamo in the
stellar envelope. The companion continues its in-spiral and eventually forms a disc which
later drives an outflow. In this case, two winds are launched from the system, both along
the polar axis. The dynamo driven outflow is expected to occur in a burst, whereas the disc
driven outflow might last ≃ 104 yrs (Blackman et al. (2001b)). Alternatively, a companion
may supply enough energy to directly unbind envelope material but subsequently shred into
a disc. The bulk of the mass is ejected along the equator while the disc material is ejected
along the rotation axis.
-- 13 --
In short, each of the three possibilities in Fig. 3 represent specific scenarios which
may occur in conjunction or individually. More work is needed to elucidate the detailed
possibilities.
4.2. Applications to specific PPNe and PNe systems
Hsia, Ip & Li (2006) took time series photometric observations of the young planetary
nebulae, Hubble 12 (HB 12; PN G111.8-02.8). The authors found evidence for an eclipsing
binary at the center in which the companion is a low-mass object (m2 < 0.443 M⊙). In
addition, there is evidence of reflection off of the secondary in the I and R bands. The
extended hourglass-like envelope of Hubble 12 suggests jet collimation. In the context of
our CE scenarios, such a collimated structure would result from either a dynamo driven
outflow in the stellar envelope or a disc driven outflow around the progenitor core. A low
enough mass binary companion is required to produce a significant polar outflow. For
α = 0.6 and a binary separation r ∼ 8 × 1010 cm, as suggested by Hsia, Ip & Li (2006), the
maximum mass that can result in a polar outflow without expelling much of the envelope
equatorially is approximately m2 ≤ 0.2M⊙. For α = 0.3, the limit is m2 ≤ 0.4 M⊙. If the
AGB progenitor were more evolved when the CE phase commenced, then the upper limit
on the corresponding masses would be lower. For instance, if the companion were immersed
in our interpulse AGB star, α = 0.6 would require m2 ≤ 0.08 M⊙ while α = 0.3 yields
m2 ≤ 0.17 M⊙. Therefore, it seems likely that the mass of the companion in Hubble 12 is
at least a factor of 2 or 3 less then the upper limit proposed by Hsia, Ip & Li (2006).
HD 44179, nicknamed the Red Rectangle, is a proto-PNe in which the secondary in
the central binary is a low-mass post-AGB star (Men'shchikov et al. (2002)). The system
consists of a disk with bipolar outflows emanating from the central region. CO maps
suggest that the circumstellar disk is in approximate Keplerian rotation (Bujarrabal et al.
(2003)). In addition, the expanion velocity in the outer region is quite low (∼ 0.4 km s−1)
suggesting that the disk is bound to the central binary. It may be that equatorial matter
ejected during a CE phase did not fully escape. The disk material then fell backwards until
the resulting angular momentum was sufficient to remain in a stable orbit.
Recent Spitzer IRS data from the Red Rectangle show evidence of oxygen-rich
material in the carbon-rich bipolar outflows (Markwick-Kemper et al. (2005)) in addition
to the oxygen-rich material in the circumbinary disc (Waters et al. (1998)). A possible
evolutionary explanation for the disc composition, is that the progenitor incurred rapid
equatorial mass loss while the star was still oxygen-rich. The carbon-rich interior layers
were exposed and used to shape the bipolar outflows. The oxygen-rich material in the
-- 14 --
carbon-rich outflows may be the result of a jet from a disk composed of a shredded brown
dwarf or planet.
NGC 7009 is a PNe, exhibiting complex morphology that includes two distinct ansae
far from the central source. Fern´andez et al. (2004) analyzed the kinematics of the ansae
and determined expansion velocities and proper motion. For a distance of 0.86 kpc to NGC
7009, the diameter of the ansae ∼ 3.8 × 1016 cm with a radial velocity of ∼ 1.3 × 107 cm/s
measured away from the nebulae. This gives an upper limit for the burst time of the ansae
(τa ≃ 100 yrs). Thus an interface dynamo operating ∼ 102 years may be responsible for the
production of ansae in some planetary nebulae.
5. Conclusions
We have studied the implications of embedding a range of low-mass (m2 < 0.3M⊙)
companions into the envelope of a 3.0 M⊙ star during three epochs of its evolution: For
RGB stars, we find that envelope ejection is unlikely. However, for AGB and interpulse
AGB stars, we find scenarios that can lead to partial or complete envelope ejection. For
an AGB star and conventional efficiency parameters (0.3 < α < 0.6), we find that massive
brown dwarfs can directly eject the envelope equatorially. Lower mass companions that do
not directly eject the envelope may spiral in far enough to induce a differential rotation
mediated dynamo that ejects material poloidally. In addition, the companion may be
shredded into a disc, possibly facilitating a disc driven outflow. For the interpulse AGB
star, the envelope has significantly expanded, further lowering the energy required to unbind
the system. In this phase, it is easiest for the envelope to be ejected.
For systems in which the envelope is directly ejected, the expected outflow is equatorial
with a torus-like appearance. The amount of envelope material contained in the outflow
is determined by the mass of the secondary and the penetration depth of the companion.
Shallow penetration depths may be indicative of higher mass companions and result in lower
tori masses and less molecule formation rates in the expanding outflow. Systems which form
discs or incur dynamos are expected to generate polar outflows. If the companion is a brown
dwarf that gets shredded inside the envelope of a carbon-rich AGB star, contamination of
the polar outflow may result in the formation of crystalline silicates or other oxygen-rich
substances.
When the dynamo occurs in the envelope, via the induced envelope differential rotation,
and there is no reseeding of this differential rotation, the outflow can only last < 100yr.
This would imply a poloidal poweful but swift jet burst (e.g. ansae) in the PPNe phase.
-- 15 --
A disc dynamo may be required for a disc mediated magnetic outflow but this would
be powered by accretion, which falls off more gradually in time. The power from a disc
mediated outflow would therefore produce an observable outflow over a longer time scale,
and into the PNe phase (Blackman et al. 2001b).
To build on the current results, more detailed calculations are needed to include the
three dimensional nature of the binary interaction, the angular distribution of the induced
outflow mass and composition, the operation of a dynamo, the inclusion of a wider range of
companion masses, and the possibility of an initally rotating envelope.
We thank A. Frank, W. Forrest, J. Kastner, and I. Minchev for useful discussions and
comments. We would also like to thank S. Kawaler for use of his evolutionary models.
JTN acknowledges financial support of a Horton Fellowship from the Laboratory for
Laser Energetics through the Department of Energy and HST grant AR-10972. EGB
acknowledges support from NSF grants AST-0406799, AST-0406823, and NASA grant
ATP04-0000-0016 (NNG05GH61G).
-- 16 --
REFERENCES
Alexander, M. E., Chau, W. Y., & Henriksen, R. N. 1976, ApJ, 204, 879
Bains, I., Gledhill, T. M., Richards, A. M. S., & Yates, J. A. 2004, in Asymmetrical
Planetary Nebulae III eds. M. Meixner, J. H. Kastner, B. Balick & N. Soker, ASP
Conf. Series, 313, 562
Balick, B., & Frank, A. 2002, ARA&A, 40, 439
Blackman, E. G., Frank, A., Markiel, J. A., Thomas, J. H. & Van Horn, H. M. 2001
Nature, 409, 485
Blackman, E. G., Frank, A., & Welch, C. 2001, ApJ, 546, 288
Blackman, E. G., Nordhaus, J. T. & Thomas, J. H. 2006 New Astronomy, 11, 452
Blackman, E. G. 2004, in ASP Conf. Ser. 313: Asymmetrical Planetary Nebulae III:
Winds, Structure and the Thunderbird, 313, 401
Bondi, H. 1952, MNRAS, 112, 195
Bujarrabal, V., Alcolea, J., & Neri, R. 1998, ApJ, 504, 915
Bujarrabal, V., Castro-Carrizo, A., Alcolea, J., & S´anchez Contreras, C. 2001, A&A, 377,
868
Bujarrabal, V., Neri, R., Alcolea, J., & Kahane, C. 2003, A&A, 409, 573
Burgasser, A. J., et al. 2002, ApJ, 564, 421
Burrows, A., Hubbard, W. B., Saumon, D., &Lunine, J. I. 1993, ApJ, 406, 158
Castro-Carrizo, A., Bujarrabal, V., S´anchez Contreras, C., Alcolea, J., & Neri, R. 2002,
A&A, 386, 633
De Marco, O., Bond, H. E., Harmer, D., & Fleming, A. J. 2004, ApJ, 602, L93
De Marco, O. & Moe, M. 2005, in Planetary Nebulae as Astronomical Tools Eds:
R. Szczerba, G. Stasinska and S. K. Gorny (AIP Conference Proceedings)
(astro-ph/0511356)
Dewi, J. D. M., & Tauris, T. M. 2000, A&A, 360, 1043
Durney, B.R., & Robinson, R.D. 1981, ApJ, 253, 290
-- 17 --
Eggleton, P. P. 1983, ApJ, 268, 368
Fern´andez, R., Monteiro, H., & Schwarz, H. E. 2004, ApJ, 603, 595
Frank, A., van der Veen, W. E. C. J., & Balick, B. 1994, A&A, 282, 554
Geballe, T. R., et al. 2002, ApJ, 564, 466
Hartigan, P., Edwards, S., & Ghandour, L. 1995, ApJ, 452, 736
Hsia, C. H., Ip, W. H. & Li, J. Z. in press, AJ 2006, astro-ph/0603224
Iben, I. J., & Livio, M. 1993, PASP, 105, 1373
Jordan, S., Werner, K., & O'Toole, S. J. 2005, A&A, 432, 273
Livio, M., & Soker, N. 1984, MNRAS, 208, 763
Markwick-Kemper, F., Green, J. D., & Peeters, E. 2005, ApJ, 628, L119
Matt, S., Frank, A., & Blackman, E. G. 2004, ASP Conf. Ser. 313: Asymmetrical Planetary
Nebulae III: Winds, Structure and the Thunderbird, 313, 449
Mauron, N., Huggins, P. J. in press AA 2006, astro-ph/0602623
Men'shchikov, A. B., Schertl, D., Tuthill, P. G., Weigelt, G., & Yungelson, L. R. 2002,
A&A, 393, 867
Morris, M. 1987, PASP, 99, 1115
Paczynski, B. 1976, IAU Symp. 73: Structure and Evolution of Close Binary Systems, 73,
75
Parker, E. N. 1993, ApJ, 408, 707
Pascoli, G. 1993, Journal of Astrophysics and Astronomy, 14, 65
Politano, M. 2004, ApJ, 604, 817
Pollard, H. 1979, Celestial Mechanics (Mathematical Association of America)
Regos, E., & Tout, C. A. 1995, MNRAS, 273, 146
Reyes-Ruiz, M., & Lopez, J. A. 1999 ApJ, 524, 952
Rudiger, G. 1989, Differential rotation and stellar convection (Berlin: Verlag)
-- 18 --
Sahai, R. 2000, ASP Conf. Ser. 199: Asymmetrical Planetary Nebulae II: From Origins to
Microstructures, 199, 209
Sahai, R. 2002, Revista Mexicana de Astronomia y Astrofisica Conference Series, 13, 133
Sandquist, E. L., Taam, R. E., Chen, X., Bodenheimer, P., & Burkert, A. 1998, ApJ, 500,
909
Shima, E., Matsuda, T., Takeda, H., & Sawada, K. 1985, MNRAS, 217, 367
Soker, N. 1997, ApJS, 112, 487
Soker, N. 2001 MNRAS, 324, 699
Soker, N. 2001, MNRAS, 324, 699
Soker, N. 2002, A&A, 386, 885
Soker, N. 2004, in Asymmetrical Planetary Nebulae III eds. M. Meixner, J. H. Kastner, B.
Balick & N. Soker, ASP Conf. Series, 313, 562
Soker, N. 2006, PASP, 118, 260
Soker, N. submitted to ApJ Letters 2006 astro-ph/0603113
Soker, N., Livio, M., & Harpaz, A. 1984, MNRAS, 210, 189
Soker, N., & Livio, M. 1994, ApJ, 421, 219
Soker, N., & Subag, E. 2005, AJ, 130, 2717
Soker, N., & Zoabi, E. 2002, MNRAS, 329, 204
Sorensen, P., & Pollacco, D. 2004, ASP Conf. Ser. 313: Asymmetrical Planetary Nebulae
III: Winds, Structure and the Thunderbird, 313, 515
Su, K. Y. L., Hrivnak, B. J., Kwok, S., & Sahai, R. 2003, AJ, 126, 848
Taam, R. E., & Sandquist, E. L. 2000, ARA&A, 38, 113
Tauris, T. M., & Dewi, J. D. M. 2001, A&A, 369, 170
Terman, J. L., & Taam, R. E. 1996, ApJ, 458, 692
Thomas, J. H., Markiel, J. A., & van Horn, H. M. 1995, ApJ, 453, 403
-- 19 --
Tout, C. A., & Pringle, J. E. 1992, MNRAS, 256, 269
Vlemmings, W. H. T., Diamond, P. J., & Imai, H. 2006, Nature, 440, 58
Waters, L. B. F. M., et al. 1998, Nature, 391, 868
Webbink, R. F. 1984, ApJ, 277, 355
Yungelson, L., Livio, M., Tutukov, A., & Kenyon, S. J. 1995, ApJ, 447, 656
Zapolski, H. S., & Salpeter, E. E. 1969, ApJ, 158, 809
This preprint was prepared with the AAS LATEX macros v4.0.
-- 20 --
Fig. 1. -- Left: Density and mass profiles for our model AGB star. The dotted line is the
core-envelope boundary. Right: Mach number and sound speed as a function of radius. The
Mach number is computed from the Keplerian motion of the planet inside the envelope.
The motion is supersonic everywhere and thus justifies our choice of accretion radius (Bondi
(1952)).
Fig. 2. -- Infall time as a function of position inside the envelope of the AGB star (left) and
interpulse AGB star (right). The solid line represents a companion of mass 0.02 M⊙ and the
dotted line is a secondary of mass 0.2 M⊙.
-- 21 --
Fig. 3. -- Three possible outcomes of our CE evolution.
(a.) The companion becomes
embedded in the stellar envelope, orbital separation is reduced, eventually resulting in
unbinding the envelope equatorially. (b.) The companion spirals in, the envelope is spun up
causing it to differentially rotate. The presence of a deep convective zone, coupled with the
differential rotation, generates a dynamo in the envelope. (c.) The companion is shredded
into an accretion disc around the core. The disc then drives an outflow which, in principle,
can unbind the envelope.
-- 22 --
Fig. 4. -- For various efficiencies α (see Eq. 1), the solid line shows the radius at which
the change in orbital energy equals the binding energy of the envelope for the beginning of
the AGB star (left) and interpulse AGB star (right). The dotted vertical line marks the
core-envelope boundary. The long-dashed line represents the radius at which the companion
is tidally shredded by the core. The short-dashed line is where the companion first fills its
Roche lobe, initiating mass transfer to the envelope.
-- 23 --
Fig. 5. -- The solid line depicts the energy required to unbind the envelope for the AGB star
(left) and interpulse AGB star (right), if the secondary is not tidally shredded as it traverses
the envelope. The dashed lines represent the amount of energy deposited into the envelope
from the change in orbital energy of the secondary for efficiency parameter α (Eq. 1). For
α = 1.0, a m2 = 0.02 M⊙ brown dwarf delivers enough energy to blow off the AGB envelope
at r ∼ 1010 cm. For α = 0.3, the brown dwarf must traverse all the way to the core-envelope
boundary before supplying enough energy to unbind the system. For smaller α, a m2 = 0.02
M⊙ companion cannot unbind the AGB envelope before spiraling down to a radius where
an interface dynamo might participate in unbinding the envelope. For the interpulse AGB
star, a 0.02 M⊙ brown dwarf can supply enough orbital energy to unbind the envelope for
α = 1.0 and α = 0.3.
-- 24 --
Fig. 6. -- Two rotation profiles for our 3.0 M⊙ AGB star. The solid curve represents the
spin up of an initial stationary envelope by an infalling 0.02 M⊙ brown dwarf. The dotted
curve is the rotation profile generated in Blackman et al. (2001a) in which a main sequence
star exhibiting solid body rotation conserves angular momentum of spherical mass shells
during its evolution onto the AGB. The solid vertical line marks the core boundary and the
short-dashed line represents the base of the convective zone. The long-dashed line is the
base of the differential rotation zone used in Blackman et al. (2001a).
|
0806.3461 | 1 | 0806 | 2008-06-20T20:07:18 | Physical approximations for the nonlinear evolution of perturbations in dark energy scenarios | [
"astro-ph",
"gr-qc",
"hep-ph"
] | The abundance and distribution of collapsed objects such as galaxy clusters will become an important tool to investigate the nature of dark energy and dark matter. Number counts of very massive objects are sensitive not only to the equation of state of dark energy, which parametrizes the smooth component of its pressure, but also to the sound speed of dark energy as well, which determines the amount of pressure in inhomogeneous and collapsed structures. Since the evolution of these structures must be followed well into the nonlinear regime, and a fully relativistic framework for this regime does not exist yet, we compare two approximate schemes: the widely used spherical collapse model, and the pseudo-Newtonian approach. We show that both approximation schemes convey identical equations for the density contrast, when the pressure perturbation of dark energy is parametrized in terms of an effective sound speed. We also make a comparison of these approximate approaches to general relativity in the linearized regime, which lends some support to the approximations. | astro-ph | astro-ph | hyi
a axiai f he iea ev i f
e bai i dak eegy
eai
. R. Aba
R. C. Baia
. ibea
ad R. Refe d
1
1
2
2
1
i de Fíi
a Uiveidade de Sã a
C 66318 05315 970 Sã a Bazi
2
i de Fíi
a Teói
a Uiveidade Ead a a ia
R. a a 145 01405 900 Sã a Bazi
E ai : abafa.if. .b baiafa.if. .b
ibeaif. e.b efe if. e.b
be 22 2018
eywd : C gy: hey (cid:22) C gy: age
a e
e f he Uivee
Aba
. The ab da
e ad diib i f
aed b je
h a ga axy
e wi be
e a ia iveigae he a e f dak eegy ad
dak ae. be
f vey aive b je
ae eiive y he
e ai f ae f dak eegy whi
h aaeize he h
e f i
e e b a he d eed f dak eegy a we whi
h deeie he
a f e e i ihgee ad
aed
e. Si
e he ev i f
hee
e be f wed we i he iea egie ad a f y e aivii
faewk f hi egie de exi ye we
ae w axiae
hee: he
wide y ed hei
a
ae de ad he e d ewia aa
h. We hw
ha bh axiai
hee
vey idei
a e ai f he deiy
a
whe he e e e bai f dak eegy i aaeized i e f a e(cid:27)e
ive
d eed. We a ake a
ai f hee axiae aa
he geea
e aiviy i he ieaized egie whi
h ed e he axiai.
8
0
0
2
n
u
J
0
2
]
h
p
-
o
r
t
s
a
[
1
v
1
6
4
3
.
6
0
8
0
:
v
i
X
r
a
hyi
a axiai f he iea ev i f e bai i dak eegy
eai2
1. d
i
We w have vewhe ig evide
e ha he Uivee i a
e eaig ib y de
he i(cid:29) e
e f e ye f egaive e e ba
e (cid:21) dak eegy DE [1 2 3℄.
weve eve h gh DE ay be die
y eib e f hi eha
ed exai i
i wide y be ieved ha he die
ia
f e bai i DE deiy ad e e
e fai i vey weak. Thi i i
y
e
y f a
gi
a
a de f DE whi
h de have e bai.
F
a a (cid:28)e d de f DE hi
e eai vey hgee eve
ga axy ad
e
a e. e ii
a y hi
a be ded a f w. hee
de he
a a (cid:28)e d
a have e axed i ii eegy ae ad e
e ie ha he ie
a e f he vaiai f he (cid:28)e d i ge ha he bb e ie
i yig a vey (cid:29)a eia . Theefe he
a a (cid:28)e d be exadiai y igh
m < H0 whee H0 i he bb e aaee day. The a f he
a a (cid:28)e d e he
a e f i aia vaiai ad he
e e a y exe
a e bai i he
a a (cid:28)e d f
a e λ < 1/m he C wave egh whi
h ae f he de f he
bb e adi . weve hi ag e ay a y e geea de f dak
eegy.
f y
e i he ev i f he ba
kg d he he e f dak eegy
i he ev i f dak ae e bai i
ee y deeied by i e ai f
ae w = pe/ρe whee pe i he hgee e e ad ρe i he hgee eegy
deiy f dak eegy [4 5 6℄. A hi eve dak eegy a(cid:27)e
e fai
idie
y be
a e a i a diae he ba
kg d vey age
e ae
ied aa by he e ig a
e eaed exai [7 8℄.
weve dak eegy
a i(cid:29) e
e
e fai i a addiia ae.
f i i a dyai
a (cid:28)e d (cid:29) id he dak eegy e ihgeeiie ad
hee e bai wi iea
gaviaia y bh wih hee ve ad wih
f dak ae [9℄. Thi ea ha e dak eegy i j a
gi
a
a
i wi bh fee ad
eae
a gaviaia eia .
A h gh he e(cid:27)e
f hee ihgeeiie i he dak eegy
e be
e
a a w → −1 i ay de wih w 6= −1 i
a be eg igib e whe ev ved
i he iea egie [10 11 12 13 14 15 16 17 18 19 20 21 22 23℄. Si
e
he e(cid:27)e
f dak eegy e bai he
i
i
wave ba
kg d ae ie
a ee e.g. [24℄
e fai i he y eaiig be f he a e f
dak eegy a ad ieediae
a e.
evehe e a f y e aivii
ehd ea iea e bai i
avai ab e. Whe hee i a e e igedie he iea e aivii
e ai ake
a vey
i
aed f. The eaîe T a Bdi TB de [25 26 27℄ i he
e e
a ge a wkig fa i b i y wk if ae i e e e.
The b e i wih he gaviy ide f he e ai b wih he iea
ev i f ae ad he e aivii
eae f e e.
hi ee
he y we died de wih ihgee dak eegy ae
hyi
a axiai f he iea ev i f e bai i dak eegy
eai3
he iv vig
ai
a
a a (cid:28)e d [15 16 17 18 19 20℄ f whi
h he e ai
f i ad he e e f w die
y f a give agagia. F hee de
he fee aaee ae he
a a eia ad e e f iiia
dii. hi
aa
h he e ai f ae he deiy e bai ad he e e e bai
ae deived aiie. e
e a e kieai
a ad de ideede aa
h
e fai
e i ii he hgee de
ii f dak eegy i
e f a aaeized e ai f ae w(z) i e y a
kig.
Thee ae w vey di(cid:27)ee axiai f b w geea e aiviy ha
have bee fe e y ed. They ae he hei
a
ae SC de [28 29 30℄ ad
he e d ewia aa
h [31 32 33 34℄. We have e
e y ed hee
axiai i he iea egie i de hw ha he (cid:16)e(cid:27)e
ive e ai f
ae(cid:17) f dak eegy iide a
aed egi
d be vey di(cid:27)ee f i ba
kg d
va e [23℄.
hi wk we hw ha eve h gh he de yig a i f eihe
aa
h ae ahe di(cid:27)ee hey yie d exa
y he ae e baive e ai
a g a he e e e bai ae eaed i he ae way. They a have
a ia advaage: hey a w f a
ee y aaeized aa
h dak
eegy. F hee we
ae he gwh f e bai i he iea egie wih
a ieaized e aivii
aa yi ad hw ha hey ae ii a edig he
axiai.
Thi ae i gaized a f w. Se
i 2 we eview bh he ad SC
aa
he ad hw ha hey ae e iva e. Se
i 3 we dy he iea ev i
f e bai i DE i hi axiai. The iea ev i f e bai i
a ivee wih a 2
e (cid:29) id i died i geea e aiviy i Se
i 4. We
ee a
ai bewee he e aivii
aa yi ad he axiae aa yi i
he iea egie i Se
i 5. Se
i 6
de.
2. Shei
a
ae ad e d ewia
gy
iea e bai hey hee ae eeia y hee degee f feed f
a a
e bai: he eegy deiy e bai δρ he e e e bai δp ad he
a a aii
e π [35 36℄. A a eaive e i give by he deiy
a
δρ/ρ he ve
iy eia θ = ~∇ · ~v ad he aii
e [37℄. Si
e age
a e
aii
ee de
ay aid y hey
a y be
e e eva agai iide
e
whi
h have
aed. Thi ea ha aii
e h d i(cid:29) e
e he a f
hee
e ad heefe i i ike y ha dak eegy de
a be di(cid:27)eeiaed
he bai f aii
e. F hi ea we d
ide i ay f he i
hi wk ee hweve [38℄.
We wi aaeize he e e e bai ig he
a ed e(cid:27)e
ive d
ve
iy [39℄ de(cid:28)ed a c2
eff ≡ δpe/δρe . We wi a e ha c2
eff i a f
i f ie
y eve h gh hi i i(cid:28)
ai a
k ay fa bai i
gi
a e bai
hey. Thi h d be
ea f he fa
ha δpe i a ideede degee f feed
hyi
a axiai f he iea ev i f e bai i dak eegy
eai4
whe ie ad aia deede
e
a be ad fe ae
ee y di(cid:27)ee f
δρe . y i a ai
a ga ge he
a ed (cid:16)e fae(cid:17) f he (cid:29) id whee T i
0 = 0
he e(cid:27)e
ive d eed
i
ide wih he ivea d eed f iea e aivii
e bai c2
X [39 40℄. ay be di(cid:30)
ea ize hi aaeizai i a a a
de b he i ai i
h di(cid:27)ee f wha hae whe we aaeize
he e ai f ae.
The ai ea ha we e he e(cid:27)e
ive d eed h gh i ha i a w
dy iea
e fai wihi he hei
a
ae de [28℄. hi
exee y i e de a hei
a y yei
egi f hgee vedeiy
ev ve iide he hgee exadig Uivee hi i he
a ed (cid:16) ha(cid:17)
deiy (cid:28) e. Geea e aivii
ag e hw ha e
a egad he vedee
egi a a ii ivee f iive
va e ad he we e he Fieda ad he
Ray
ha dh y e ai ev ve he deiy ad adi f he hei
a egi [29 30℄.
i heefe exee y ieeig ha hi i i(cid:28)ed e aivii
aa
h
i
ide wih a e d ewia aa
h
gy. fa
we wi hw be w
ha a g a he e e e bai ae de
ibed i e f a e(cid:27)e
ive e e
he w axiai ae
ee y e iva e. Thi ea ha he ai hyi
a
haa
eii
f gaviaia
ae f
e
h a ga axy
e i bab y
we de
ibed wihi hi faewk.
The ag e i a f w. Fi he SC aa
h h d be a gd axiai
f age
a e whee e aivii
e(cid:27)e
h d ae b e
eai y f
a
a e whee he (cid:16)ii ivee(cid:17) ag e i e e aive. he he had
he aa
h i we ivaed by he hyi
f gaviy i a
a e b i
a ed wk f age
a e. Tha he w aa
he
i
ide hw ha a ea i
e iied ee he e ai f he SC/ aa
h h d give a gd de
ii
f he gaviaia iea
i
a e a e ha he bb e adi .
2.1. e d ewia
gy
gy ai
e i a
vig gid aa
ea
h he gaviaia y wih
a ewia eia . The ii f he ai
e i he gid ae he e bed
vaiab e. A h gh bvi y iied hi aa
h
a be ed f ay
(cid:28)g ai
y he hei
a y yei
e. B i de big he aa
h
e
he SC de we wi ad he ae bai
a i f he SC de f he
gi
a e bai.
We
ide a adix e f w (cid:29) id
d dak ae ad dak eegy. The key
a i f he SC de ee he ex be
i ae ha he deiy f ea
h (cid:29) id
i hgee a a ie i he hei
a egi hi i he ha deiy (cid:28) e
ad ha he ve
iy (cid:28) e eeve hi hgeeiy.
The
vig
diae ae ~x0 = ~r0/a whee ~r0
i he hgee
e bed hyi
a dia
e (cid:21) hee he adi f a hei
a y yei
egi.
Ude he a i f he SC de he e bed hyi
a dia
e hyi
a adi
hyi
a axiai f he iea ev i f e bai i dak eegy
eai5
a be wie a:
~r = [a (t) + f (t, ~x0)] ~x0 ,
1
whee a i he a
a e fa
ad f i he f
i ha a
f he deviai
f he ba
kg d ev i. The hyi
a ve
iy i he give by:
~u =
d~r
dt
=(cid:16) a + f(cid:17) ~x0 = H +
f
a! ~r0 ,
2
whee = ∂/∂t ad H = a/a i he bb e aaee. F he a e a iy we
a
de(cid:28)e a e(cid:27)e
ive ae f exai f he hei
a egi:
h = H +
f
a
.
Si
e he e bed ve
iy i e aed he e
ia ve
iy ~v by
~u = a ~x0 + ~v ,
we bai f E. 2 ha:
~v = f ~x0 .
ai
a he divege
e f hi ve
iy (cid:28)e d i give by:
θ ≡ ~∇ · ~v = 3 f + ~x0 · ~∇ f .
3
4
5
6
B f a ha (cid:28) e he a e vaihe ad we bai a i e e ai bewee
he
a exai ae h ad he ba
kg d exai ae H :
h = H +
f
a
= H +
θ
3a
.
The
gi
a de i de
ibed by he e ai [31℄:
+ ~∇ · (~uj ρj) + pj ~∇ · ~uj = 0 ,
~∇pj
ρj + pj
,
∂ρj
∂t
∂~uj
∂t
+(cid:16)~uj · ~∇(cid:17) ~uj = −~∇Φ −
(ρk + 3pk) ,
∇2Φ = 4πGXk
7
8
9
10
whee ρj pj ad ~uj dee ee
ive y he deiy e e ve
iy f a give
i
(cid:29) id ad Φ i he ewia gaviaia eia d e a he
e; he
e ai ae wie i hyi
a
diae. The
edig e bai abve
he ba
kg d ae deed by δρj δpj ~vj ad φ. Thee e ai ae ee
ive y
geea izai f (cid:29) id wih e e f he
i iy e ai f he E e e ai
bh va id f ea
h (cid:29) id e
ie j ad f he i e ai. i
e he abe
e f
a e ai ha di
ae he ev i f e e: i hi hyddyai
a aa
h
e e i a hedyai
a f
i f he eegy eea e e
.
F
d dak ae ad bay he e e i ze b f dak eegy hee
i a hgee a we a a ihgee e e. The hgee e e
hyi
a axiai f he iea ev i f e bai i dak eegy
eai6
i a y de
ibed i e f a aaeized e ai f ae we(t)
h ha
pe(t) = we(t)ρe(t). A f he e e e bai we have
he e
ify ahe
fee f
i he e(cid:27)e
ive d eed c2
effδρe . Wihi he SC de
ii
hi ea ha we
ide a e(cid:27)e
ive e ai f ae wc iide he hei
a egi
eff δpe = c2
whi
h i e
eai y e a he ba
kg d e ai f ae.
Wih he a i f he SC de he e ai f
gy a e a
i e f. Uig he deiy
a δj ≡ δρj/ρj we bai afe e a geba:
θj
eff j − wj(cid:1) δj +
δj + 3H(cid:0)c2
θj + Hθj +
a (cid:2)1 + wj +(cid:0)1 + c2
eff j(cid:1) δj(cid:3) = 0 ,
ρ0 kδk(cid:0)1 + 3c2
eff k(cid:1) .
θ2
j
3a
= −4πGaXk
11
12
E. 11 f w f he
i iy e ai ad E. 12 i he divege
e f he
E e e ai. The a e a iy i E. 12 i f d by ig he i e ai.
e ha i geea we have eaae E e e ai f ea
h (cid:29) id [21℄ b f a
ha (cid:28) e
~∇δj = 0 hey be idei
a hee i y e θ. The
ea f ha i bvi : i de eeve he ha (cid:28) e a (cid:29) id (cid:29)w i
he ae way. e
e i hi axiai we have ehig ii a a e(cid:27)e
ive
ig e (cid:29) id de
ii [41℄.
2.2. The hei
a
ae de
e w bie(cid:29)y eview he hei
a
ae de . Thi fa i de
ibe a
hei
a y yei
egi f if eegy deiy ρc = ρ0 + δρ ieed i a
hgee ivee f eegy deiy ρ0 . Thi hei
a egi wi dea
h f he
exai f he Uivee ad eve a y
ae.
Cide he
i iy e ai f ea
h (cid:29) id deed by a idex j i he hei
a
egi:
ρcj + 3h(cid:0)1 + wcj(cid:1) ρcj = 0 ,
13
whee h = r/r i he
a exai ae f ha egi ad wcj dee he e ai
f ae i he e bed egi. We
a egad hi hei
a egi a a Fieda
Uivee wih aia
va e [28℄. The dyai
f he
diae r i he give by
he e
d Fieda e ai a ied hi
aig egi:
r
r
= −
4πG
3 Xj (cid:0)ρcj + 3pcj(cid:1) .
14
E ai 13 ad 14 whi
h wee baied ig geea e aivii
ag e ae
he bai
e ai f he SC de . e ha hee i y e dyai
a e ai
f he
aig egi whi
h i i ageee wih he ig e E e e ai ha we
f d f he ve
iy (cid:28)e d i he de
ii E. 12.
The e e ad he eegy deiy ide he hei
a egi ae e aed by he
ba
kg d e ai f ae p0j = w0j ρ0j . ide he hei
a egi hee aiie
hyi
a axiai f he iea ev i f e bai i dak eegy
eai7
a be di(cid:27)ee f hei ba
kg d va e we have w pcj = wcj ρcj f he
aig egi. de
ae he SC fa i wih he e ai deived
i he a e
i we wi e y hee he ae e(cid:27)e
ive d eed we ed befe
i de de
ibe he e e e bai. e
e we eed exe he e ai
f ae wcj i e f c2
eff j . Uig he deiy
a δj = δρj/ρ0j we have ha:
ρcj = (1 + δj) ρ0j ,
f whi
h i f w ha:
wcj =
pcj
ρcj
=
p0j + δpj
ρ0j + δρj
= wj +(cid:0)c2
eff j − wj(cid:1)
δj
1 + δj
.
15
16
Thi e ai e ae he e ai f ae i he e bed egi he ba
kg d
e ai f ae he e(cid:27)e
ive d eed ad he ize f e bai. i ib e
ha he a e f dak eegy
a be igi(cid:28)
a y
haged i he e bed egi a
hee we d bbed (cid:16)dak eegy ai(cid:17) [23℄. Thi e(cid:27)e
i geea
ig
eve a he eve f iea ev i ad i agi de deed he dyai
a
ev i f D ad DE (cid:29)
ai (cid:21) ee a Ref. [16 20℄.
Uig w E. 15 ad 16 we
a e
a E. 13 a:
We
a e iiae h ig E. 7 wih he e :
δj + (3h − 3H) (1 + wj) (1 + δj) + 3h(cid:0)c2
δj + 3H(cid:0)c2
eff j − wj(cid:1) δj +(cid:2)1 + wj +(cid:0)1 + c2
eff j − wj(cid:1) δj = 0 .
eff j(cid:1) δj(cid:3)
w
ide he dyai
a e ai 14. F E. 7 we
a wie:
θ
a
= 0 .
17
18
h =
r
r
− h2 = H +
θ
3a
− H
θ
3a
,
ad bi ig hi exei i E.14 we bai wih he he f E. 15 16
ha:
θ + Hθ +
θ2
3a
= −4πGaXk
ρ0k δk(cid:0)1 + 3c2
eff k(cid:1) .
19
E ai 17 ad 19 ae idei
a E.
11 12. Thi ea ha bh
aa
he ae idei
a . The e ai 7 ad 16 eab e a ae he
vaiab e i he SC vaiab e ad w i be
e
ea ha he w di(cid:27)ee
de
ii give he ae dyai
f a ha e bai whee e e gadie
ae abe.
3. iea ev i i he SC/ aa
h
Eve h gh we hwed ha he ad SC aa
he ae e iva e ha i de
ea ha hey ae
e
. Uf ae y ee y hee i f y iea
geea eae f he ev i f e bai i Geea Re aiviy GR. F hi
ea we wi
ae ieaized e wih he baied f ieaized GR.
hyi
a axiai f he iea ev i f e bai i dak eegy
eai8
We wi
e he iea ev i f a vedee egi we iide he ae
diaed ea ad wi
ae he gwig de baied i he /SC fa i
wih he e aivii
gwig de.
The (cid:28) de e ai
a be ieaized ad e
a a a ig e e
d de
di(cid:27)eeia e ai f he deiy
a f ea
h (cid:29) id e
ie. We wi a e ha
hee i a way a dia d ad a bdia s (cid:29) id. Uig he
a e fa
a f
he ie ev i
′ = d/da we bai f he dia e
ie:
δ′
d
whee
+
3
2
3δd
δ′′
d +
(1 − wd)(cid:21)
a (cid:20)3∆d +
2a2 (cid:2)∆d (1 − 3wd) − (1 + wd)(cid:0)1 + 3c2
∆d =(cid:0)c2
eff d − wd(cid:1) .
3δd
2a2 = 0 ,
+
δ′′
d +
3
2
δ′
d
a
eff d(cid:1)(cid:3) = 0 .
F
d dak ae ceff = w = 0 hi e ai ed
e
20
21
22
24
25
26
27
28
wih he we kw gwig i δ(a) ∝ a. e
e whe
d dak ae i
dia whi
h h d be he
ae i he iea egie he iea ev i f i
deiy e bai i he adad e.
F he e geea
ae f a dia (cid:29) id wih
a e ai f ae ad
a eed f d he i i give by:
δd (a) = c1a1+3wd + c2a−3(1+2c2
eff d
−wd)/2.
23
T ig w he he b dia (cid:29) id i e bai bey he e ai:
δ′′
s +
δ′
s
3
2
a (cid:20)3∆s +
3δs
2a2 [∆s (1 − 3wd)] =
(1 − wd)(cid:21)
3δd
+
whee
2a2 (1 + ws)(cid:0)1 + 3c2
eff d(cid:1) ,
A ig agai ha ws ad c2
∆s =(cid:0)c2
eff s − ws(cid:1) .
eff s ae
a e ha he i:
δs (a) = c3a−3∆s + c4a(3ws−1)/2 + c5a1+3wd ,
whee
c5 =
1
2
c1 (1 + ws)(cid:0)1 + 3c2
eff d(cid:1)
(1 + wd)(cid:19) +
×(cid:20)(1 + 3wd)(cid:18)∆s +
1
2
1
2
(1 − 3wd) ∆s(cid:21)−1
aie f a ai
a i f he ihgee e ai.
ai
a if ae i he dia (cid:29) id i f w ha:
c5 =
c1 (1 + ws)
3∆s + 1
hyi
a axiai f he iea ev i f e bai i dak eegy
eai9
ad he dak eegy deiy
a gw i he ae way a he dak ae deiy
a.
addii f he
ae i whi
h c2
eff = w e ha a adiabai
dii
ai(cid:28)ed ae y δe = (1 + we)δm . geea hweve he e bai have a
adiabai
e ad he dak eegy deiy
a ev ve a:
δs (a) =
(1 + ws)
3∆s + 1
δd (a) + c3a−3∆s ,
29
whee he a e i he igh had ide i a de
eaig de i
ae.
4. iea Ev i i GR
a evi ae [21℄ we hwed ha f a ig e efe
(cid:29) id wih e e
gadie he gwig de i he ieaized SC/ aa
h
i
ide wih he
f d wih Geea Re aiviy GR. w we wa
ae he ad he GR
i f he dak eegy e bai i he iea egie i
dig e e
gadie. We wi
ide hee e bai d ig he ae diaed eid
i.e. whi e DE i bdia. Thi i ivaed by he fa
ha beved
e wee fed we i he ae diaed eid
We
ide
a a e bai he ei
i he ewia ga ge wih
aii
e:
ds2 = (1 + 2φ) dt2 − a2 (1 − 2φ) d~x2 .
The (00) ad (ii)
ee f Eiei e ai i F ie a
e ae:
k2
δρj ,
a2 φ + 3H(cid:16) φ + Hφ(cid:17) = −4πGXj
+ H 2(cid:19) φ = 4πGXj
φ + 4H φ +(cid:18)2
a
a
δpj ,
ad he
evai e ai T µ
0;µ = 0 ad T µ
i;µ = 0 yie d:
δj + 3H(cid:0)c2
θj + H(cid:0)1 − 3c2
eff j − wj(cid:1) δj + (1 + wj)(cid:18)θj
sj(cid:1) θj −
(1 + wj) ρja
k2δpj
a
−
whee c2
sj = pj/ ρj i he adiabai
eed f d. .
− 3 φ(cid:19) = 0 ,
k2
a
φ = 0 ,
30
31
32
33
34
ay he ev i f e bai i a ye
iig f dak eegy
ad dak ae i ieaized GR i de
ibed by he f wig e f 5
ed
di(cid:27)eeia e ai:
φ + 4H φ +(cid:18)2
a
a
+ H 2(cid:19) φ =
3
2
H 2Ωec2
effδe ,
δm +
θm
a
− 3 φ = 0 ,
35
36
hyi
a axiai f he iea ev i f e bai i dak eegy
eai10
δe + (1 + we)(cid:18)θe
a
θm + Hθm −
k2
a
φ = 0 ,
− 3 φ(cid:19) + 3H(cid:0)c2
eff − we(cid:1) δe = 0 ,
θe + H(cid:0)1 − 3c2
s e(cid:1) θe −
k2c2
effδe
(1 + we) a
−
k2
a
φ = 0 .
5. Cai bewee GR ad
37
38
39
gy he iea ev i f D ad DE i deeied by he ye f
e ai ha aie f E. 8 ad 9 f ea
h (cid:29) id ae y:
δm +
θm
a
= 0 ,
δe + (1 + we)
θm + Hθm −
θe
a
k2
a
k2c2
+ 3H(cid:0)c2
eff − we(cid:1) δe = 0 ,
φ = 0 ,
θe + Hθe −
effδe
(1 + we) a
−
k2
a
φ = 0 .
40
41
42
43
i
e he abe
e f a dyai
a e ai f φ. T e iiae he k2φ e
we
a e he
ai i ied by he i e ai i
gy E.10.
The he ie vaiai f he eia i deeied by he ev i f he deiy
a. A i
e ha hee e ai a
k e e whe
aed wih hei
e aivii
ea a a eady ied i Ref. [42℄. weve a we wi hw
d ig he ae diaed ea ad a
a e hi di
ea
y
hage y he
ve
iy ha DE e bai de
ay b d dify i ae ie behavi .
he ae diaed egie φ = const. i a i f E.32 whi
h a
aie f he ye f E.40 43.
hi
ae i i ieeig i
e ha
E.37 be
e idei
a E.41. weve E.39
i
ide wih E.43 y i
he
ae cs e = 0.
A we ee he e ai f he gwh f e bai ae di(cid:27)ee i GR ad
a eady i he iea egie. w we ef a aiaive dy f hi di(cid:27)ee
e.
We wi wk he
ae f a ae diaed ivee wih a a DE
e
a exe
ed i he iea egie i whi
h
ae φ i a
a. F hee avid
f he
i
ai we a e
a va e f w ad c2
eff .
Ude hee
dii we
a wie a e
d de di(cid:27)eeia e ai f he
iea gwh f he dak eegy deiy e bai a a f
i f he
a e fa
δe(a):
δ′′
e + α
δ′
e
a
+(cid:20)β +
k2c2
eff
a2H 2(cid:21) δe
a2 = − (1 + w)
k2
a2H 2
φ
a2 .
44
hyi
a axiai f he iea ev i f e bai i dak eegy
eai11
Thi e ai aie bh i GR ad : i he ae
ae we kee he e e gadie
i he E e e ai 9 whi
h wa ded i he
ae f a ha e bai. y
he aaee α ad β ae di(cid:27)ee i he w
ae:
αGR =
αP N =
3
2
3
2
+ 3∆ − 3w ; βGR = 3∆(cid:18)1
2
− 3w(cid:19)
+ 3∆ ; βP N =
3
2
∆ ,
whee ∆ wa de(cid:28)ed i E. 25.
45
46
We wi ake a
ai f
ig a
a e whee he axiai
i ed be e a
ae. hi
ae we
a eg e
he β e i he ae
ba
ke f E.44 ad we iediae y wie a
a ai
a i:
δe = −
(1 + w)
c2
eff
φ.
47
de ve he hgee e ai we ef he f wig
hage f
vaiab e:
δe (a) = x1−αy (x) ,
48
whee x i de(cid:28)ed i e f he
fa ie η a x = kceffη . The E.44 be
e:
d2y
dx2 +
1
x
dy
dx
+(cid:20)1 −
µ2
x2(cid:21) y = 0 ,
whee a
dig he di(cid:27)ee
e(cid:30)
ie i E.
va e:
49
45 46 µ a e di(cid:27)ee
µGR = ±
µP N = ±
50
51
1
1
2(cid:0)1 − 6c2
eff(cid:1) ,
2(cid:0)1 + 6w − 6c2
eff(cid:1) .
The i ae Bee f
i f (cid:28) kid J±µ (x). The dak eegy deiy
a behave a:
δe (x) = x1−αJ±µ (x) −
(1 + w)
c2
eff
φ.
52
Thee i bh have a
i ay behavi wih a de
eaig a i de
ia x1−α−1/2
− (1 + w) φ/c2
eff .
ad hey eve a y ea
h he
a va e δe =
de
he
k hi aa yi
a behavi we ei
a y ve he
ee ye
eff = −we = 0.8
m = 0.75 ad we ev ved he e ai f a iiia edhif f zi = 100.
f
ed di(cid:27)eeia e ai 35 39. We ed a i ai c2
Ω(0)
de = 1 − Ω(0)
We exaied he de k = 100H0 = 0.0236hMpc−1
f λ = 266h−1Mpc we iide bb e adi a zi ad age e gh be i he iea
φi = 0
a f he ae we e
. A iiia
dii we
he φi = −10−4
edig a hyi
a
a e
ad:
δm(zi) = −2φi(cid:20)1 +
k2(1 + zi)2
3H(zi)2 (cid:21) ; δe(zi) = (1 + w)δm(zi) ;
53
hyi
a axiai f he iea ev i f e bai i dak eegy
eai12
θm(zi) =
2(1 + zi)k2
3H(zi)
φi ; θe(zi) = 0 ,
54
whi
h ae
ie wih Eiei e ai ad adiabai
iy. The e i eeed
i Fig. 1 ad
aed he de
ay fa
ad he (cid:28)a va e give by E.47.
-2
-3
-4
-5
-6
-7
e
d
∆
0
1
g
o
L
0.25
0.5
0.75
1
1.25
1.5
1.75
2
Log101+z
Fig e 1. iea ev i f dak eegy e bai i GR a a
a e
k = 0.0236hMpc−1
f c2
eff = −we = 0.8. The id ie i he i f he
ee e f 5
ed di(cid:27)eeia e ai. The ded ie i he de
ay fa
a
dig E.52. The dahed ie i he ai
a i E. 47.
We a ef he ae exe
ie f he axiai. The ei
a
i f he ye f E. 40 43 wih he ae aaee ad iiia
dii
i eeed i Fig. 2 ad
aed he de
ay fa
ad he (cid:28)a va e give by
E.47. Agai we ee ha he a iaive aa yi
a behavi i ed
ed by he
ei
a i.
Theefe eve h gh he e ai f GR ad ae he ae a eady a
he iea eve he e ae a iaive y di(cid:27)ee. ai
a bh aa
he
edi
he ae ayi
behavi f he DE e bai. Be
a e he de
ay ae
f he aie i igh y di(cid:27)ee i ea
h
ae he ie whe he ayi
egie
i ea
hed di(cid:27)e (cid:21) i he aa
h hi hae a a ae ie.
A hi i we h d
a aei he igi f a aae di
ea
y
bewee he e baied i hi Se
i ae y a
a behavi f he DE
e bai E. 47 ad he e baied i Se
i 3 whee we hwed ha
i he SC/ aa
h he DE e bai gw a D e bai E. 29.
The ea i ha i Se
i 3 we a ed a ha (cid:28) e f he e bai whi
h
a eig kceff = 0 i he ae ba
ke f E.
44.
hi
ae he
ai
a i i
δe(a) = −
1 + w
β
k2φ
H 2a2 =
3(1 + w)
2β
δm(a) ,
55
hyi
a axiai f he iea ev i f e bai i dak eegy
eai13
-2
-3
-4
-5
-6
-7
e
d
∆
0
1
g
o
L
0.25
0.5
0.75
1
1.25
1.5
1.75
2
Log101+z
Fig e 2. iea ev i f dak eegy e bai i he axiai
a a
a e k = 0.0236hMpc−1
f c2
eff = −we = 0.8. The id ie i he
i f he
ee e f 4
ed di(cid:27)eeia e ai. The ded ie i
he de
ay fa
a
dig E.52. The dahed ie i he ai
a i
E. 47.
whee i he a e a iy we ed i e ai f he
ae f a dia dak
ae
e k2φ = −(3/2)H 2a2δm . e
e we ee ha ideed i hi
ae i
fa
f e bai wih a a de be k he e bai i DE gw a
he ae a
e a he D e bai i he iea egie.
e bb e
a e
gy i exe
ed be va id d e i
ihee y iaee iea
i:
ideed i ha faewk e bai wih
a e age ha he bb e adi w d behave i he ae way a he we iide i.
weve i
e we ae y ieeed i he ev i f e bai whi
h ae iiia y
i he iea egie ad we iide he bb e adi hi ia
h i ie eva.
Theefe we d
ae he e bai wih he GR e bai i age
a e.
A a (cid:28)a eak we e
a ha he aa yi
i E.52 i va id y f
iea e bai d ig he ae diaed eid.
hi egie he ae
deiy
a gw a δm ∝ a ad φ i
a i ie. Whe he
e ee
he iea egie ae (cid:29)
ai gw fae i.e δm ∝ an
wih n > 1
he he gaviaia eia h d gw i ie. e
e DE behavi i iea
e i exe
ed be di(cid:27)ee f he iea aa yi e . weve he
ayi
a i i E.52 i va id d ig he iiia iea
e f
ae
ae ad DE (cid:29)
ai
a gw wih φ.
6. C
i
The dy f e bai i dak eegy ha e
eived a gea dea f aei
e
e y. DE e bai have he eia a e he
e f age
a e
e
hyi
a axiai f he iea ev i f e bai i dak eegy
eai14
fai i he ivee. The exie
e f DE e bai
a i i
i e be eed
i f e vey
h a he Dak Eegy S vey DES [43℄ ad EUCD [44℄. Thee
f e bevai ay he diig ih ag di(cid:27)ee de f DE.
S
e fai
d ig he iea age f he ev i f
e bai. Uf ae y hee i ig aa yi
a de
ii f hi iea
age i f GR. Axiai ehd be ed. ib y he ed ehd
i bdy i ai b d e i vey ieive
ig e iee i i
a
i
a whe e wa dy di(cid:27)ee de . F hee bdy i ai
e y ewia hyi
ad d a w f he ibi iy f DE (cid:29)
ai.
hi ae we dy w di(cid:27)ee axiai
hee ae y he Shei
a
C ae ad e d ewia aa
he. The advaage f hee
hee i ha
DE
a be f y
haa
eized by 2 f
i: he e ai f ae aaee w(z) ad
he e(cid:27)e
ive eed f d ceff(z). We hw ha de a iia e f a i
i i ib e a ae e aa
h i he he edeig he
ee y
e iva e. de
ae hee axiai wih GR we dy e bai
i he ieaized egie wih a aa
he. Whe he a i ab he e e
e bai ae he ae bh i GR ad /SC we (cid:28)d ha he (cid:29)
a
i ee
he ae a iaive behavi edig he axiai. weve i
de eab ih e (cid:28) y he va idiy f he axiai i he iea egie
a
ai h d be ade wih e e baive de i GR
h a a
exeded TB
a f de i
dig (cid:29) id wih e e. Wk a g hi die
i
i i ge.
Refee
e
[1℄ . Fiea . T e ad D. ee 2008 aXiv:0803.0982 [a h℄.
[2℄ A. A be
h e a . 2006 a h/0609591.
[3℄ U. Se jak A. S a ad .
Da d CA 0610 014 2006 a h/0604335.
[4℄ . . E. eeb e ad B. Raa Rev. d. hy. 75 559 2003 a h/0207347.
[5℄ T. adaabha hy. Re. 406 49 2005 g
/0311036.
[6℄ V. Sahi e
. e hy. 653 141 2004 a h/0403324.
[7℄ E. V. ide hy. Rev. D72 043529 2005 a h/0507263.
[8℄ . ibea ad R. Refe d CA 0607 009 2006 a h/0604071.
[9℄ . Cb e S. Dde ad . A. Fiea hy. Rev. D55 1851 1997 a h/9608122.
[10℄ S. Bai ak Ahy. . 590 636 2003 a h/0303112.
[11℄ A. yi A. V. a
i R. aiii ad S. A. Be Ahy. . 599 31 2003 a
h/0303304.
[12℄ E. V. ide ad A. eki . . Ry. A. S
. 346 573 2003 a h/0305286.
[13℄ E. . ka . Bde ad Y. (cid:27)a . . Ry. A. S
. 349 595 2004 a
h/0309485.
[14℄ . h e . E. Sigai A. R. Zee . S. B
k ad . R. ia
k . . Ry. A.
S
. 357 387 2005 a h/0402210.
[15℄ D. F. a ad C. va de B
k A. Ahy. 421 71 2004 a h/0401504.
[16℄ . . e ad D. F. a . . Ry. A. S
. 368 751 2006 a h/0409481.
[17℄ . . e A. C. da Si va ad . Aghai A. Ah. 450 899 2006 a h/0506043.
[18℄ C. e ad . Bege . . Ry. A. S
. 360 1393 2005 a h/0504465.
hyi
a axiai f he iea ev i f e bai i dak eegy
eai15
[19℄ . aea ad D. F. a . . Ry. A. S
. 371 1373 2006 a h/0504519.
[20℄ S. D a ad . a hy. Rev. D75 063507 2007 g
/0612027.
[21℄ . R. Aba R. C. Baia . ibea ad R. Refe d
CA 0711 012 2007
aXiv:0707.2882 [a h℄.
[22℄ D. F. a D. . Shaw ad . Si k 2007 aXiv:0709.2227 [a h℄.
[23℄ . R. W. Aba R. C. Baia . ibea ad R. Refe d hy. Rev. D77 067301 2008
aXiv:0710.2368 [a h℄.
[24℄ . R. Aba F. Fie i ad T. S. eeia hy. Rev. D70 063517 2004 a h/0405041.
[25℄ G. eaie Ge. Re . Gav. 29 641 1997.
[26℄ R. C. T a
. a. A
ad. S
i. 20 169 1934.
[27℄ . Bdi . . Ry. A. S
. 107 410 1947.
[28℄ . E. G ad . G . Ri
had Ahy. . 176 1 1972.
[29℄ T. adaabha S
e Fai i he Uivee Cabidge Uiveiy e 1993.
[30℄ . Fa ba ad E. Gazaaga . . Ry. A. S
. 301 503 1998 a h/9712095.
[31℄ . A. S. ia V. Za
hi ad R. . Badebege . . Ry. A. S
. 291 1 1997
a h/9612166.
[32℄ .
. wag ad . h 1997 a h/9701137.
[33℄ R. R. R. Rei hy. Rev. D67 087301 2003.
[34℄ .
. wag ad . h Ge. Re . Gav. 38 703 2006 a h/0512636.
[35℄ . . Badee hy. Rev. D22 1882 1980.
[36℄ . daa ad . Saaki g. The. hy. S . 78 1 1984.
[37℄ C. . a ad E. Be
hige Ahy. . 455 7 1995 a h/9506072.
[38℄ T. ivi ad D. F. a 2007 0707.0279.
[39℄ W. Ahy. . 506 485 1998 a h/9801234.
[40℄ V. khav hyi
a F dai f C gy Cabidge Uiveiy e 2005.
[41℄ . . Ave i . . G. Be
a ad C. . A. . ai hy. Rev. D77 101302 2008 0802.0174.
[42℄ R. aiii 2008 0806.0516.
[43℄ Dak Eegy S vey T. Abb e a . 2005 a h/0510346.
[44℄ The EUCD ea 2008 h://
i.ea.i/
ie
e e/www/aea/idex.
f?faeaid=102.
|
0708.2947 | 1 | 0708 | 2007-08-22T01:45:42 | Detection of high-energy gamma rays from winter thunderclouds | [
"astro-ph",
"physics.geo-ph"
] | A report is made on a comprehensive observation of a burst-like $\gamma$-ray emission from thunderclouds on the Sea of Japan, during strong thunderstorms on 2007 January 6. The detected emission, lasting for $\sim$40 seconds, preceded cloud-to-ground lightning discharges. The burst spectrum, extending to 10 MeV, can be interpreted as consisting of bremsstrahlung photons originating from relativistic electrons. This ground-based observation provides first clear evidence that strong electric fields in thunderclouds can continuously accelerate electrons beyond 10 MeV prior to lightning discharges. | astro-ph | astro-ph | Detection of high-energy gamma rays from winter thunderclouds
H. Tsuchiya,1 T. Enoto,2 S. Yamada,2 T. Yuasa,2 M. Kawaharada,1 T. Kitaguchi,2
M. Kokubun,3 H. Kato,1 M. Okano,1 S. Nakamura,4 and K. Makishima1, 2
1Cosmic Radiation Laboratory, Riken, 2-1,
Hirosawa, Wako, Saitama 351-0198, Japan
2Department of Physics, University of Tokyo,
7-3-1, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
3Department of High Energy Astrophysics,
Institute of Space and Astronautical Science, JAXA,
3-1-1, Sagamihara, Kanagawa 229-8501, Japan
4Department of Physics, Tokyo University of Science,
1-3, Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
(Dated: May 8, 2014)
Abstract
A report is made on a comprehensive observation of a burst-like γ-ray emission from thunder-
clouds on the Sea of Japan, during strong thunderstorms on 2007 January 6. The detected emission,
lasting for ∼40 seconds, preceded cloud-to-ground lightning discharges. The burst spectrum, ex-
tending to 10 MeV, can be interpreted as consisting of bremsstrahlung photons originating from
relativistic electrons. This ground-based observation provides first clear evidence that strong elec-
tric fields in thunderclouds can continuously accelerate electrons beyond 10 MeV prior to lightning
discharges.
PACS numbers: 82.33.Xj, 92.60.Pw, 93.85.-q
7
0
0
2
g
u
A
2
2
]
h
p
-
o
r
t
s
a
[
1
v
7
4
9
2
.
8
0
7
0
:
v
i
X
r
a
1
INTRODUCTION
As first suggested by Wilson [1], intense electric fields in thunderclouds and lightning are
considered to accelerate charged particles to relativistic energies, which in turn will produce
Bremsstrahlung photons. Thus, detailed investigation of such photons is expected to provide
a valuable key to the particle acceleration process in strong electric fields.
Thunderclouds are found to have electric fields reaching 100 kVm−1 [2], or sometimes
even > 400 kVm−1 [3]. Usually, frequent collisions in the dense atmosphere will not allow
electrons to be accelerated even under such a strong electric field. However, seed electrons
more energetic than ∼ 100 keV, if once produced by, e.g., cosmic rays, can gain energies
from the electric field fast enough to overcome the collisional energy losses. Thus, such
electrons are accelerated and multiplied in the atmosphere, because their mean free paths
get progressively longer, up to ∼ 50 m, as they acquire higher energies. This is the concept
of runaway electron avalanche mechanism [4]. When these conditions are fulfilled, electrons
can be accelerated in thunderclouds to several tens MeV energies or higher, and then produce
Bremsstrahlung γ-rays.
Actually, sudden non-thermal x-/γ-ray emissions from upper atmosphere, called terres-
trial γ-ray flashes, have been observed at equatorial latitudes by near Earth satellites [5, 6].
The spectra of the flashes are roughly expressed by power-law functions with the photon
indices of ∼ −1 [7], some of them extending up to 10 − 20 MeV. Now they have been
interpreted as arising from high-altitude electrical discharges above thunderstorms. In addi-
tion, ground-based observatories have also detected similar photon bursts from natural [8, 9]
and rocket-triggered [10, 11] lightning, revealing that the bursts are mostly associated with
lightning discharge processes. The burst spectra in the rocket triggered lightning mostly
extend up to ∼ 250 keV, or to 10 MeV [12] on rare occasions. These bursts including the
terrestrial γ-ray flashes have typical duration of milliseconds or less.
In addition to those short-duration bursts, more prolonged radiation from thunder ac-
tivity, lasting for a few tens seconds to minutes, have been observed by detectors at high
mountains during summer seasons [13, 14]. Also radiation-monitoring posts, arranged in
nuclear power plants in the coastal area of Sea of Japan, have frequently detected such
prolonged events associated with winter thunderstorms [15]. However, compared with the
short duration events, the longer duration ones have remained much less understood, due
2
to inadequate information on, e.g., the particle species, their arrival direction and energy
spectrum. Here we show one observation revealing those essential features of the prolonged
burst.
EXPERIMENT
Kashiwazaki-Kariwa nuclear power plant is located in the coastal area of Sea of Japan
in Niigata prefecture, where thunder activity is very high in November through January.
Installed at the rooftop of a building in this power plant, our new automated radiation de-
tection system has been continuously and successfully operated since 2006 December 22. The
system consists of two independent and complementary parts, Detector-A and Detector-B,
placed with a separation of ∼ 10 m. Designed after the Hard X-ray Detector experiment [16]
on board the 5th Japanese cosmic x-ray satellite Suzaku, Detector-A consists of two iden-
tical main detectors, each having a structure shown in Fig. 1; it comprises a cylindrical
NaI scintillator operated in 40 keV − 3 MeV, surrounded by a well-type BGO (Bi4Ge3O12)
scintillator shield counter of which the average thickness is around 2.0 cm. Placed above the
two main counters is a 5 mm thick plastic scintillator. Detector-A records the energy deposit
of each event, and its arrival time with a time resolution of 10 µs. As an important feature
of Detector-A, the surrounding BGO shields prevent low-energy environmental radiations
from impinging on the NaI scintillator. Furthermore, using the signals of the BGO shields,
we can roughly know where an incident particle comes from. When the NaI scintillator is
triggered by an incident particle and the NaI signals does not coincide with the BGO ones,
the incident particle would generally arrive from the vertical direction neither the horizontal
one nor the ground since the BGO shields firmly cover side and bottom of each NaI scin-
tillator. Thus, using the BGO signals in anti-coincidence, we can reduce background of the
NaI scintillator, and give it broad directional sensitivity towards the vertical direction.
Aiming at higher energy ranges, Detector-B utilizes NaI and CsI scintillation counters
observing in 40 keV − 10 MeV and 600 keV − 80 MeV, respectively. With a spherical
shape, both work as omni-directional monitors without anti-coincidence shields. Each of
them accumulates pulse-height spectra in 1024 channels every 6 seconds, and record broad-
band pulse counts every second.
In addition to Detectors A and B, we measure environmental light and electric field. Three
3
light sensors are used to measure optical intensity in different directions, each consisting of a
Si photo-diode (HAMAMATSU S1226-8BK) and a handmade analog circuit. Output signals
of each light sensor are fed to a 12 bit ADC, and recorded as ADC values every 0.1 sec. The
electric field sensor is a commercial product (BOLTEK EFM-100), and its analog output is
collected by a 12 bit ADC every 1 sec and converted to the electric field strength between
± 10 kVm−1 with a resolution of 5 Vm−1. Details of our system are reported by Enoto [17],
and will be published elsewhere.
RESULTS
On 2007 January 6, the coastal area of Sea of Japan was covered with one of the strongest
thunderstorms in this winter. Figure 2 shows count rate histories of the BGO and NaI
scintillators of Detector-A, and of the NaI and CsI counters of Detector-B, obtained from
21:10 to 22:10 UT on this day (6:10 to 7:10 local standard time on January 7). Superposed
on gradual background changes due to fall-outs of radioactive Radon, we have detected a
remarkable count enhancement at about 21:43, lasting for ∼ 1 min. This event, or ”burst”,
is thus highly significant in all the inorganic scintillators of our system.
Figure 3 gives details of the count-rate histories between 21:40 and 21:50 UT. Records
of light and electric fields are also shown. The signal in Detector-A is most prominent in
the > 3 MeV range covered by the last overflow bin [Fig . 3(a)], where the excess counts
in a 36 sec of 21:43:09 − 21:43:45 UT, calculated by subtracting the background from the
total counts during the burst, become 129 ± 17 and 43.6 ± 3.5, without and with the BGO
anticoincidence, respectively. The NaI and CsI counters of Detector-B recorded 3 − 10 MeV
excess counts of 74.2 ± 9.6 and 53.3 ± 8.4, respectively [Fig. 3(b)]. Therefore, the burst
intensity from the two Detectors (without anticoincidence for A) agrees with one another,
since the values from Detector-A sum the two counters.
At 21:44:18 UT, or 70 seconds after the burst onset, the optical data rapidly increased
for about 1 second [Fig. 3(c)], with a rise time less than the sampling time which is 100 ms.
In coincidence with this flash, the electric filed quickly changed its polarity from positive
values (≥ 5 kVm−1) during the burst to negative (≤ −10 kVm−1), although the signal is
saturated at ±10 kVm−1 due to the limited range of the electric field sensor. Thus, the
first flash can be interpreted as a lightning discharge; so are the subsequent four flashes.
4
The burst had already ceased by the first lightning, and no increase accompanied any of the
total five discharges. Thus, the observed event appears to be associated with thunderclouds
themselves, rather than to the lightning discharges.
During the burst, the plastic scintillator of Detector-A [Fig. 3(e)], with its lower threshold
set at 1 MeV, showed no significant excess signals. Quantitatively, the number of events
detected by the plastic scintillator with an area of 464 cm2 in the 36 sec is 185 ± 14, which
is consistent with the number, 191, predicted by its longer-term average count. The lack of
excess signals in the plastic scintillator, together with the long-burst duration, excludes the
burst being due to electrical noise. Furthermore, we can conclude that the burst is dominated
by γ-rays, since the thin plastic scintillator has high detection efficiency for charged particles
but is nearly transparent to γ-rays. Another important information, derived from the data,
is that the photons came from a sky, probably covered with thunderclouds, rather than
from the ground. Actually, excess count ratio between NaI [Fig.2(b)] and BGO [Fig.2(a)] of
Detector-A, ∼ 0.18 ± 0.03, is significantly higher than their normal count ratio, 0.08, which
is due mostly to environmental radioactivity.
Figure 4 shows pulse-height spectra of the burst from Detector-A and B, accumulated
for the same 36 sec; we subtracted background averaged over 21:30 − 21:40 UT and 21:45
− 21:55 UT. Thus, the excess counts in both systems exhibit very hard continuum spectra.
The spectra of Detector-B, in particular, extend up to 10 MeV. Supposing that the γ-rays
came from the thunderclouds, these spectra imply that electrons were accelerated in the
thunderclouds at least to 10 MeV, or to higher energies in order to efficiently produce the
10 MeV photons.
Based on results from an experiment [12], and from a Monte Carlo simulation [18], we may
assume that the photon number spectrum arriving at the top of our system has a power-law
form as α × Eβ, where α is normalization in units of photons MeV−1cm−2, β is the photon
index, and E is the photon energy in MeV. In this work, the data points in 600 keV − 10
MeV where both NaI and CsI counters can cover are selected [Fig. 4(b)]. Then, convolving
this model with the detector responses, and fitting it to the observed spectra (but excluding
data points with too large errors), we obtained α = 2.63 ± 0.92 and β = −1.58 ± 0.23
(χ2/d.o.f. = 3.3/3) from NaI of Detector-B, while α = 1.93 ± 0.42 and β = −1.69 ± 0.15
(χ2/d.o.f. = 5.3/6) from CsI. Thus, the two sets of results are consistent. Weighted means of
the two estimates become α = 2.04±0.38 and β = −1.66±0.13. The spectra from Detector-A
5
are generally consistent with this. In reality, the observed photons must be produced at some
distance, and propagate in the atmosphere while they are attenuated mainly via Compton
scattering. When corrected for the mild energy dependence of the Compton cross section,
the photon index softens to −2.01 ± 0.14, −2.44 ± 0.15, and −3.03 ± 0.16, assuming the
source distance of 300, 600, and 1,000 m, respectively.
DISCUSSIONS
With respect to the duration and its precedence to the lightning, the present event is
similar to one event observed by Torii et al. [15], on another nuclear plant located on the
same coastal area with a separation of 300 km. However, compared with their results, the
present work provides a few new findings. That is, we have confirmed that the signal is
dominated by γ-rays arriving from the sky direction. In addition, our results reveal that the
photon spectrum extends up to 10 MeV.
The photon spectrum detected in the present work, extending up to 10 MeV, clearly
indicates a non-thermal emission mechanism. The most promising candidate process is
Bremsstrahlung. Then, we can draw a following possible scenario to explain the present
observation [17]. Generally, a winter thundercloud along the Sea of Japan is considered
to have 3-layer (positive-negative-positive) charge structure in the vertical direction [19].
Then, the lower part of such thunderclouds is positively charged. Electrons, first produced
by cosmic rays, will be accelerated by the runaway electron avalanche process [4] from the
middle part to the cloud bottom (or top), emitting Bremsstrahlung γ-rays toward the ground
(or sky). Since the γ-rays must be relativistically beamed into a narrow cone (∼ 3◦ half-cone
angle for a 10 MeV electron), we can detect such photons only under rare conditions when
the cone happens to fall on our detector.
Let us briefly evaluate energetics of the above scenario. Since we observed a fluence of
1 − 10 MeV photons as np = 2.4 × 104 m−2, the total number of these photons at the source
Np becomes d2npǫ−1∆Ω, where d is the source distance, ǫ is the atmospheric attenuation
factor, and ∆Ω is a solid angle of the possibly beamed emission. The values of ǫ for 3 MeV
photons are calculated, e.g., as 0.3, 0.08, and 0.01 for d = 300, 600, and 1000 m, respectively.
Employing ∆Ω = 8.6 × 10−3 (or equivalently a half-cone angle of ∼ 3◦) which is appropriate
for 10 MeV electrons, we obtain Np ∼ 108(d/300m)2(0.3/ǫ). Since a 10 MeV electron
6
traveling ∼ 50 m in the atmosphere is expected to lose a minor fraction, η ∼ 10−2 [20], of its
energy into 3 MeV bremsstrahlung photons, the overall energy of the produced relativistic
electrons is estimated as
Ee = Np/η × 10 MeV ∼ 10−2(d/300m)2(0.3/ǫ) J.
(1)
Of course, the estimate is subject to rather large uncertainties, because we may not have
observed the peak of the radiation beam. Nevertheless, the estimated Ee is a negligible
fraction of an energy associated with a single lightning, 107 − 1010 J [21], ensuring that the
scenario is energetically feasible.
In summary, we have successfully shown that electrons are accelerated beyond 10 MeV
in thunderclouds, for an extended period, prior to lightning discharges rather than in co-
incidence with them. Thus, this kind of observation of non-thermal x-/γ-ray emissions
associated with thunderclouds and lightning will provide a valuable key to the particle ac-
celeration processes in strong electric fields.
T. Torii kindly gave us helpful advice on our experiment. We greatly thank members
of radiation safety group of Kashiwazaki-Kariwa power station, Tokyo Electric Power Com-
pany, including K. Oshimi, T. Nakamura, K. Inomata, and Y. Iwasaki, for support of our
experiment. We are grateful to S. Otsuka and Y. Ikegami for design and construction of
housings of our system. This work is supported in part by the Special Research Project for
Basic Science in RIKEN, titled “Investigation of Spontaneously Evolving Systems”. The
work is also supported in part by Grant-in-Aid for Scientific Research (S), No.18104004.
7
[1] C. T. R. Wilson, Proc. of Phys. Soc. London 37, 32 (1925).
[2] T. C. Marshall, W. D. Rust, and M. Stolzenburg, J. Geophys. Res. 100, 1001 (1995).
[3] T. C. Marshall et al., Geophys. Res. Lett. 32, L03813 (2005).
[4] A. V. Gurevich, G. M. Milikh, and R. Roussel-Dupre, Phys. Lett. A 165, 463 (1992).
[5] G. J. Fishman et al., Science 264, 1313 (1994).
[6] D. M. Smith et al., Science 307, 1085 (2005).
[7] R. J. Nemiroff, J. T. Bonnell, and J. P. Norris, J. Geophys. Res. 102, 9659 (1997).
[8] C. B. Moore et al., Geophys. Res. Lett. 28, 2141 (2001).
[9] J. R. Dwyer et al., Geophys. Res. Lett. 32, L01803 (2005).
[10] J. R. Dwyer et al., Science 299, 694 (2003).
[11] J. R. Dwyer et al., Geophys. Res. Lett. 31, L05118 (2004).
[12] J. R. Dwyer et al., Geophys. Res. Lett. 31, L05119 (2004).
[13] A. P. Chubenko et al., Phys. Lett. A 275, 90 (2000).
[14] Y. Muraki et al., Phys. Rev. D69, 123010 (2004).
[15] T. Torii, M. Takeishi, and T. Hosono, J. Geophys. Res. 107, 4324 (2002).
[16] T. Takahashi et al., Publ. Astron. Soc. Japan 59, S35 (2007).
[17] T. Enoto, Master Thesis, University of Tokyo (in Japanese) (2007).
[18] T. Torii et al., Geophys. Res. Lett. 31, L05113 (2004).
[19] N. Kitagawa and K. Michimoto, J. Geophys. Res. 99, 10713 (1994).
[20] B. Rossi, High-Energy Particles (Prentice-Hall Inc., Englewood Cliffs, N.J., 1965), p. 51.
[21] T. C. Marshall and M. Stolzenburg, J. Geophys. Res. 106, 4757 (2001).
8
(cid:22)(cid:15)(cid:17)
(cid:20)(cid:17)(cid:22)(cid:15)(cid:17)
(cid:81)(cid:77)(cid:66)(cid:84)(cid:85)(cid:74)(cid:68)(cid:1)(cid:84)(cid:68)(cid:74)(cid:79)(cid:85)(cid:74)(cid:77)(cid:77)(cid:66)(cid:85)(cid:80)(cid:83)
(cid:49)(cid:46)(cid:53)
(cid:49)(cid:46)(cid:53)
(cid:25)(cid:20)(cid:15)(cid:19)
(cid:47)(cid:66)(cid:42)(cid:1)(cid:9)(cid:53)(cid:77)(cid:10)
(cid:25)(cid:20)(cid:15)(cid:19)
(cid:18)(cid:17)(cid:18)(cid:15)(cid:23)
(cid:19)(cid:22)(cid:15)(cid:21)
(cid:35)(cid:40)(cid:48)
(cid:18)(cid:18)(cid:26)(cid:15)(cid:21)
(cid:49)(cid:46)(cid:53)
(cid:49)(cid:46)(cid:53)
(cid:49)(cid:46)(cid:53)
(cid:18)(cid:17)(cid:23)(cid:15)(cid:24)
(cid:18)(cid:19)(cid:15)(cid:24)
(cid:18)(cid:18)(cid:26)(cid:15)(cid:21)
(cid:25)(cid:20)(cid:15)(cid:19)
(cid:18)(cid:18)(cid:26)(cid:15)(cid:21)
(cid:52)(cid:74)(cid:69)(cid:70)(cid:1)(cid:55)(cid:74)(cid:70)(cid:88)
(cid:53)(cid:80)(cid:81)(cid:1)(cid:55)(cid:74)(cid:70)(cid:88)
FIG. 1: Schematic views of the two main sensors and the plastic scintillator of Detector-A. PMT
means ”photomultiplier”. Numbers represent scales in millimeter.
9
)
1
-
c
e
s
0
2
(
s
t
n
u
o
C
)
1
-
c
e
s
0
2
(
s
t
n
u
o
C
)
1
-
c
e
s
0
2
(
s
t
n
u
o
C
)
1
-
c
e
s
0
2
(
s
t
n
u
o
C
(a)
(b)
38000
36000
34000
3500
3000
4000 (c)
3800
(d)
3600
3400
900
800
700
21:20
21:30
21:40
21:50
22:00
UT
FIG. 2: Count rates per 20 seconds of 4 inorganic scintillators between 21:10 and 22:10 UT on
2007 January 6. All abscissa represent universal time, and all errors are statistical 1σ. Panels (a)
and (b) show the count rates in > 40 keV data from the BGO and NaI scintillators of Detector-A,
respectively. Panels (c) and (d) give > 40 keV and > 600 keV data from the NaI and CsI counters
of Detector-B, respectively. The gaps in panels (a) and (b) are due to regular interruption of data
acquisition of Detector-A every hour.
10
)
)
1
1
c
c
e
e
s
s
6
6
(
(
s
s
t
t
n
n
u
u
o
o
C
C
)
1
-
c
e
s
6
(
s
t
n
u
o
C
C
C
D
D
A
A
l
l
a
a
c
c
i
i
t
t
p
p
O
O
)
)
1
1
-
-
m
m
V
V
k
k
(
(
E
E
)
1
-
c
e
s
6
(
s
t
n
u
o
C
80
80 (a)
70
70
60
60
50
50
40
40
30
30
20
20
10
10
0
0
25 (b)
20
15
10
5
0
3500 (c)
3500
3000
3000
2500
2500
2000
2000
1500
1500
1000
1000
500
500
10
10 (d)
5
5
0
0
-5
-5
-10
-10
50 (e)
45
40
35
30
25
20
15
21:40 21:41 21:42 21:43 21:44 21:45 21:46 21:47 21:48 21:49 21:50
UT
FIG. 3: Detailed count rate histories of the radiation sensors, and the light and electric field
monitors, between 21:40 and 21:50 UT. All abscissa are universal time. All bins except those in
panels (c) and (d) have 1σ statistical errors. (a) Count histories per 6 seconds in > 3 MeV energies
from the main NaI sensors of Detector-A, without (black) and with (red) the BGO anticoincidence.
(b) 3 − 10 MeV counts every 6 seconds from the NaI (black) and CsI (red) scintillators of Detector-
B. (c) One-second optical data variations. (d) One-second electric field data variations. (e) > 1
MeV counts every 6 seconds from the plastic scintillator of Detector-A.
11
410
310
210
)
1
-
V
e
M
(
s
t
n
u
o
C
(a)
(b)
10
1
-110
-210
-110
1
Energy (MeV)
-110
-110
1
1
10
10
Energy (MeV)
Energy (MeV)
FIG. 4: Background-subtracted photon energy spectra accumulated over the 36 sec. All horizontal
and vertical axes show the photon energy and the counts per each energy bin, respectively. Error
bars are statistical 1σ. (a) Spectra from Detector-A. Black and red circles show data of the main
NaI counters (summed over the two units), without and with the BGO anticoincidence. (b) Spectra
from Detector-B. Black and red squares are from NaI and CsI counters, respectively. Black and
red dashed lines show predictions of the best-fit incident power-law model, determined over the
energy range where the dashed lines are drawn.
12
|
astro-ph/0702712 | 1 | 0702 | 2007-02-27T13:37:45 | A multiwavelength study of the ultracompact HII region associated with IRAS 20178+4046 | [
"astro-ph"
] | We present a multiwavelength study of the ultra compact HII region associated with IRAS 20178+4046. This enables us to probe the different components associated with this massive star forming region. The radio emission from the ionized gas was mapped at 610 and 1280 MHz using the Giant Metrewave Radio Telescope (GMRT), India. We have used 2MASS $J H K_{s}$ data to study the nature of the embedded sources associated with IRAS 20178+4046. Submillimetre emission from the cold dust at 450 and 850 $\mu$m was studied using JCMT-SCUBA. The high-resolution radio continuum maps at 610 and 1280 MHz display compact spherical morphology. The spectral type of the exciting source is estimated to be $\sim$ B0.5 from the radio flux densities. However, the near-infrared (NIR) data suggest the presence of several massive stars (spectral type earlier than O9) within the compact ionized region. Submillimetre emission shows the presence of two dense cloud cores which are probably at different evolutionary stages. The total mass of the cloud is estimated to be $\sim$ 700 -- 1500 $\rm M_{\odot}$ from the submillimetre emission at 450 and 850 $\mu$m. The multiwavelength study of this star forming complex reveals an interesting scenario where we see the presence of different evolutionary stages in star formation. The ultra compact HII region coinciding with the southern cloud core is at a later stage of evolution compared to the northern core which is likely to be a candidate protocluster. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. 6462ms
September 5, 2018
c(cid:13) ESO 2018
7
0
0
2
b
e
F
7
2
1
v
2
1
7
2
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A multiwavelength study of the ultracompact HII region
associated with IRAS 20178+4046
A. Tej1, S. K. Ghosh1, V. K. Kulkarni2, D. K. Ojha1, R. P. Verma1 and S. Vig1 ⋆
1 Tata Institute of Fundamental Research, Homi Bhabha Road, Colaba, Mumbai 400 005, India
2 National Centre for Radio Astrophysics, Post Bag 3, Ganeshkhind, Pune 411 007, India
Received xxx / Accepted yyy
ABSTRACT
Aims. We present a multiwavelength study of the ultra compact HII region associated with IRAS 20178+4046. This
enables us to probe the different components associated with this massive star forming region.
Methods. The radio emission from the ionized gas was mapped at 610 and 1280 MHz using the Giant Metrewave
Radio Telescope (GMRT), India. We have used 2MASS JHKs data to study the nature of the embedded sources
associated with IRAS 20178+4046. Submillimetre emission from the cold dust at 450 and 850 µm was studied using
JCMT-SCUBA.
Results. The high-resolution radio continuum maps at 610 and 1280 MHz display compact spherical morphology. The
spectral type of the exciting source is estimated to be ∼ B0.5 from the radio flux densities. However, the near-infrared
(NIR) data suggest the presence of several massive stars (spectral type earlier than O9) within the compact ionized
region. Submillimetre emission shows the presence of two dense cloud cores which are probably at different evolutionary
stages. The total mass of the cloud is estimated to be ∼ 700 -- 1500 M⊙ from the submillimetre emission at 450 and
850 µm.
Conclusions. The multiwavelength study of this star forming complex reveals an interesting scenario where we see the
presence of different evolutionary stages in star formation. The ultra compact HII region coinciding with the southern
cloud core is at a later stage of evolution compared to the northern core which is likely to be a candidate protocluster.
Key words. infrared: ISM -- radio continuum: ISM -- ISM: H II regions -- ISM: individual objects: IRAS 20178+4046
1. Introduction
The early stages and evolution of massive stars are
some of the least understood aspects of star formation.
Ultracompact (UC) HII regions are manifestations of newly
formed massive (O or early B) stars deeply embedded in
the parental cloud. IRAS 20178+4046 (G78.44+2.66) is a
massive star forming region chosen from the catalog of mas-
sive young stellar objects by Chan et al. (1996). It has been
classified as an UC HII region (Kurtz et al. 1994;1999). The
kinematic distance estimates range from 1.5 kpc to 3.3 kpc
(Kurtz et al. 1994; Caswell et al. 1975). In this paper we
assume a distance of 3.3 kpc. The far-infrared (FIR) lumi-
nosity from the IRAS fluxes is estimated to be 7.0 × 104 L⊙
(Kurtz et al. 1994). IRAS low-resolution spectra show a
red continuum beyond ∼ 13 µm with the presence of the
polycyclic aromatic hydrocarbon (PAH) feature at 11.3 µm
(Volk & Cohen 1989). In their study of UC HII regions,
Faison et al. (1998), present the intermediate resolution in-
frared spectra of IRAS 20178+4046, where PAH features
at 3.3, 8.7 and 11.3 µm, the [Ne II] line at 12.8 µm and the
silicate feature at 9.7 µm have been detected. MSX mid-
infrared (MIR) image at 21 µm is presented by Crowther &
Conti (2003).
Send offprint requests to: A. Tej, email: [email protected]
⋆ Present address: INAF-Osservatorio Astrofisico di Arcetri,
Largo E. Fermi, 51-50125, Italy
In a recent study, Verma et al. (2003) have mapped this
source in two FIR bands (∼ 150 and 210 µm) with the 1 m
TIFR balloon-borne telescope. They also present the ISO
observations using the ISOCAM instrument in seven spec-
tral bands (3.30, 3.72, 6.00, 6.75, 7.75, 9.62 and 11.4 µm).
The ISOCAM images mainly consist of a single component
with a lobe towards south. They have modelled this source
with a single core and a dust density distribution of the
form r−2 comprising mainly of silicates. IRAS 20178+4046
has been part of several maser surveys (Codella et al. 1996;
Baudry et al. 1997; Slysh et al. 1999; Szymczak et al. 2000).
No maser has been found to be associated with this source.
It has been detected in the CS(2 -- 1) survey of Bronfman
et al. (1996). McCutcheon et al. (1991) and Wilking et al.
(1989) have observed this source in the CO lines as part of
the survey of protostellar candidates.
In this paper, we present a multiwavelength study of
this UC HII region. In Sect. 2, we describe the radio contin-
uum observations and the related data reduction procedure
used. In Sect. 3, we discuss other available datasets used in
the present study. Section 4 presents a comprehensive dis-
cussion of the results obtained. In Sect. 5, we discuss the
general star forming scenario in the region associated with
IRAS 20178+4046 and in Sect. 6 we summarize the results.
2
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
2. Observations and Data Reduction
2.1. Radio continuum observations
In order to probe the ionized gas component, radio contin-
uum interferometric mapping of the region around IRAS
20178+4046 was carried out using the Giant Metrewave
Radio Telescope (GMRT) array, India. Observations were
carried out at 1280 and 610 MHz. The GMRT has a "Y"-
shaped hybrid configuration of 30 antennas, each 45 m in
diameter. There are six antennas along each of the three
arms (with arm length of ∼ 14 km). These provide high
angular resolution (longest baseline ∼ 25 km). The rest of
the twelve antennas are located in a random and compact
arrangement within 1 × 1 km2 near the centre and is sen-
sitive to large scale diffuse emission (shortest baseline ∼
100 m). Details of the GMRT antennas and their config-
urations can be found in Swarup et al. (1991). The radio
sources 3C48 and 3C286 were used as the primary flux cal-
ibrators, while 2052+365 was used as phase calibrator for
both the 1280 and 610 MHz observations.
Data were reduced using AIPS. The data sets were care-
fully checked using tasks UVPLT and VPLOT for bad data
(owing to dead antennas, bad baselines, interference, spikes,
etc). Subsequent editing was carried out using the tasks
UVFLG and TVFLG. Maps of the field were generated by
Fourier inversion and subsequent cleaning using the task
IMAGR. Several iterations of self calibration were carried
out to obtain improved maps.
3. Other available datasets
3.1. Near-infrared data from 2MASS
Near-infrared (NIR) (JHKs) data for point sources around
IRAS 20178+4046 have been obtained from the Two
Micron All Sky Survey1 (2MASS) Point Source Catalog
(PSC). Source selection was based on the 'read-flag' which
gives the uncertainties in the magnitudes. In our sample
we retain only those sources for which the 'read-flag' values
are 1 -- 3. The 2MASS data have been used to study the
sources associated with the UC HII region.
3.2. Mid-infrared data from MSX
The Midcourse Space Experiment2(MSX) surveyed the
Galactic plane in four MIR bands centered at 8.3 (A), 12.13
(C), 14.65 (D) and 21.34 µm (E) at a spatial resolution of
∼ 18′′ (Price et al. 2001). The MIR data is used to model
the continuum SED from the interstellar medium around
IRAS 20178+4046. Other than this, two of the MSX bands
(A & C) cover the unidentified infrared bands (UIBs) at
6.2, 7.7, 8.7, 11.3 and 12.7 µm. The MSX images in the four
1 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 NASA
and the NSF.
2 This research made use of data products from the Midcourse
Space Experiment. Processing of the data was funded by the
Ballistic Missile Defense Organization with additional support
from NASA Office of Space Science. This research has also made
use of the NASA/ IPAC Infrared Science Archive, which is oper-
ated by the Jet Propulsion Laboratory, Caltech, under contract
with the NASA.
bands for the region around IRAS 20178+4046 have been
used to study the emission from the UIBs and to estimate
the spatial distribution of temperature and optical depth of
the warm interstellar dust. The MSX Point Source Catalog
(MSX PSC) is used to study MIR sources associated with
the UC HII region.
3.3. Sub-mm data from JCMT
Submillimetre observations at 450 and 850 µm using the
Submillimetre Common-User Bolometer Array (SCUBA)
instrument of the James Clerk Maxwell Telescope3 (JCMT)
were retrieved from the JCMT archives and processed us-
ing their standard pipeline SCUBA User Reduction Facility
(SURF). JCMT-SCUBA observations for the data used in
our study were carried out on 17 Aug 1998. Uranus was
used as the primary flux calibrator for the maps. The sub-
millimetre maps were used to study the cold dust environ-
ment associated with IRAS 20178+4046.
4. Results and Discussion
4.1. Radio continuum emission from ionized gas
High-resolution radio maps were generated from the GMRT
data. The radio continuum emission from the ionized gas as-
sociated with the region around IRAS 20178+4046 at 1280
and 610 MHz is shown in Fig. 1. The details of the obser-
vations and maps are given in Table 1. For the 610 MHz
observations, we present the map generated by using only
baselines larger than 1 kilo-lambda. Correction factors for
the system temperature (Tsys) were obtained from addi-
tional observations of the neighbouring sky on and off the
Galactic plane. The correction factors are 1.2 and 1.6 for
1280 and 610 MHz, respectively. The generated maps were
scaled using the Tsys correction factors obtained from our
observations. The integrated flux densities from our maps
at 1280 and 610 MHz are 57±4 and 66±4 mJy, respectively.
The flux densities are estimated by integrating down to
the lowest selected contour at the 3σ level, where σ-s are
the rms noises in the respective maps. The corresponding
size of the UC HII region is ∼ 10′′and 15′′at 1280 and 610
MHz, respectively. Considering the errors, these flux den-
sities are consistent with the hypothesis that the emission
at these frequencies are optically thin. For this UC HII re-
gion, Wilking et al. (1989) estimate flux density values of
80 and 65 mJy from their 15 and 5 GHz radio maps and
McCutcheon et al. (1991) obtain a value of 69 mJy at 5
GHz. As expected from our low frequency GMRT data, the
5 GHz flux density values of the above studies also indicate
an optically thin continuum flux distribution. The higher
value at 15 GHz from Wilking et al. (1989) may be due to
the contribution from the extended emission.
Kurtz et al. (1994;1999) have also studied this source at
8.3 and 15 GHz. Kurtz et al. (1994) estimate flux density
values of 82.2 and 35.6 mJy from their very high-resolution
(< 1′′) maps using the VLA B array at 8.3 and 15 GHz,
respectively. The integration boxes used by them are 9′′.4 ×
3 This paper makes use of data from the James Clerk
Maxwell Telescope Archive. The JCMT is operated by the Joint
Astronomy Centre on behalf of the UK particle Physics and
Astronomy Research Council, the National Research Council of
Canada and the Netherlands Organisation for Pure Research.
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
3
)
0
0
0
2
J
(
I
N
O
T
A
N
L
C
E
D
I
40 58 00
57 45
30
15
00
56 45
30
15
00
55 45
30
20 19 44
42
9
2
1
8
4
6
5
7
3
40
RIGHT ASCENSION (J2000)
38
36
40 58 00
57 45
30
15
00
56 45
30
15
00
55 45
30
)
0
0
0
2
J
(
I
N
O
T
A
N
L
C
E
D
I
34
32
20 19 44
42
Cont peak flux = 2.2271E-02 JY/BEAM
Levs = 1.400E-03 * (3, 5, 9, 13, 15)
Cont peak flux = 2.8967E-02 JY/BEAM
Levs = 1.460E-03 * (3, 5, 9, 13, 15, 19)
40
RIGHT ASCENSION (J2000)
38
36
34
32
Fig. 1. High-resolution radio continuum maps at 1280 MHz (left) and 610 MHz (right) for the region around IRAS
20178+4046. The rms noises are ∼ 1.4 and 1.5 mJy/beam and the synthesized beam sizes are 4′′.9 × 2′′.7 and 5′′.0×4′′.8
at 1280 and 610 MHz, respectively. 2MASS sources with spectral type estimations earlier than B0.5 (see Sect. 4.2) are
marked on the 1280 MHz map. The plus sign marks the position of the IRAS point source.
Table 1. Details of the radio interferometric continuum
observations of the ionized region associated with IRAS
20178+4046.
Details
Date of Obs.
Primary beam
Cont. bandwidth (MHz)
Synth. beam
Position angle. (deg)
Peak Flux (mJy/beam)
rms noise (σ)(mJy/beam)
Largest detectable
scale size
Highest angular
resolution
Int. Flux density‡ (mJy)
1280 MHz
2 Aug 2003
26′
16
610 MHz
18 Sep 2004
54′
16
4′′.9 × 2′′.7
5′′.0×4′′.8†
-38.6
22.3
1.4
∼ 8′
∼ 2′′
57±4
-0.25
29.0
1.5
∼ 17′
∼ 4′′
66±4
† For 610 MHz, the synthesized beam size is for the baselines >
1 kilo-lambda
‡ Flux densities are obtained by integrating down to the contour
at 3σ level, where σ represents the map noise in corresponding
frequency band.
8′′.9 and 9′′.5×8′′.5 for the two bands, respectively. Further
observations by Kurtz et al. (1999) at 8.3 GHz with the
VLA D array give a value of 68 and 158 mJy for 50′′ and
300′′ boxes, respectively with a synthesized beam size of
8′′.7×7′′.8. It should be noted here that the 300′′ box covers
a much larger area and includes the surrounding extended
diffuse emission.
The contour maps at 1280 and 610 MHz generated from
the GMRT observations display a simple compact spherical
morphology. In comparison, the high-resolution (< 1′′) ra-
dio map at 8.3 GHz of Kurtz et al. (1994) show a cometary
UC HII region. However, in their 15 GHz map, the sources
are resolved out and morphology looks 'clumpy'. This is
possibly the reason for the low integrated flux density (35.6
mJy) obtained at this frequency. The lower resolution (8′′.7
× 7′′.8) 8.3 GHz map by Kurtz et al. (1999) is unresolved
with an extended east-west emission seen to the south
which according to the authors is unlikely to be connected
with the UC HII region. A point worth mentioning here is
that we possibly detect extended diffuse emission in the low
resolution map at 610 MHz (not presented in this paper).
But the signal-to-noise ratio of the map is not sufficient
to conclusively comment on the nature and association if
any of this extended diffuse emission with the UC HII re-
gion. However, we do not detect any such emission in our
1280 MHz map. This could be because the flux levels of
the extended emission is below our sensitivity limit at 1280
MHz.
Using the low-frequency flux densities at 1280 and 610
MHz from our GMRT observations and the 8.3 GHz data
from Kurtz et al. (1999), we derived the physical properties
of the compact core of the HII region associated with IRAS
20178+4046. Mezger & Henderson (1967) have shown that
for a homogeneous and spherically symmetric core, the flux
density can be written as
S = 3.07 × 10−2Teν2Ω(1 − e−τ (ν))
(1)
4
where
τ (ν) = 1.643a × 105ν−2.1(EM )T −1.35
e
(2)
where S is the integrated flux density in Jy, Te the elec-
tron temperature in K, ν the frequency of observation in
MHz, τ the optical depth, Ω the solid angle subtended by
the source in steradians, and EM the emission measure in
cm−6pc. Also, a is a correction factor and we used a value
of 0.99 (using Table 6 of Mezger & Henderson 1967) for
the frequency range 0.6 -- 8 GHz and Te = 8000K. The two
GMRT maps were convolved to a common angular resolu-
tion of 8′′.7 × 7′′.8, which is the resolution of the 8.3 GHz
map of Kurtz et al. (1999). In our case, since the core is
unresolved, Ω is taken as this synthesized convolved beam
size (i.e Ω = 1.133 × θa × θb, where θa and θb are the half
power beam sizes). The peak flux densities of the core in
the 0.6 -- 8.3 GHz frequency range are consistent with an
optically thin HII region. Assuming a typical electron tem-
perature of 8000 K, these flux densities were used to fit the
above equations. The best fit value for the emission mea-
sure is 2.9 ± 0.6 × 105cm−6pc. We obtained an estimate of
1.4 × 103cm−3 for the electron density: ne = (EM/r)0.5,
with r being the core size which in this case corresponds to
the synthesized convolved beam size of 8′′.7 × 7′′.8. These
values are on the lower side compared to the estimates
of Kurtz et al. (1994) obtained from their high-resolution
maps at 8.3 GHz. They derive values of 2 × 106cm−6pc and
5.2 × 103cm−3 for EM and ne, respectively. It should be
noted here that the beam size of Kurtz et al. (1994) is much
smaller compared to the convolved beam size used in our
derivation.
Taking the total integrated flux densities of 57 and 66
mJy estimated from the 1280 and 610 MHz maps and using
the formulation of Schraml & Mezger (1969) and the table
from Panagia (1973; Table II), we estimate the exciting zero
age main-sequence star (ZAMS) of this UC HII region to
be of spectral type B0.5 -- B0. This is consistent with the
estimates of Kurtz et al. (1994; 1999). However, the FIR
luminosity of 7.0 × 104L⊙ from IRAS PSC implies an excit-
ing star of spectral type O8 (Kurtz et al. 1994). Radiative
transfer modeling of the FIR and MIR flux densities of
IRAS 20178+4046 by Verma et al. (2003) also suggests an
O7 ZAMS exciting star. In a later section (see Sect. 4.4), a
detailed model of the emission from the ISM around IRAS
20178+4046 has been presented, which implies the region
to be powered by a cluster with the most massive star being
of type B0.
4.2. Embedded cluster in the near-infrared
The 2MASS Ks - band image of the region around IRAS
20178+4046 is shown in Fig. 2. We see the presence of a
stellar group/cluster mostly concentrated around the UC
HII region. We use the 2MASS JHKs data to study the
nature of these sources seen in the vicinity of this UC HII
region associated with IRAS 20178+4046. We select a large
area of 90′′ radius centred on the IRAS point source so as
to completely cover the UC HII region and the surround-
ing diffuse nebulosity seen in the near-infrared images. We
have restricted our sample of sources to those having good
quality photometry (2MASS 'read-flag' 1 -- 3). Figure 3
shows the two ((J − H)/J and (H − K)/K) NIR colour-
magnitude (CM) diagrams. Using the ZAMS loci and the
reddening vectors, we estimated the spectral type of the
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
o
+ 4 0 5 9 ' 0 0 "
3 0 "
o
+ 4 0 5 8 ' 0 0 "
3 0 "
o
+ 4 0 5 7 ' 0 0 "
3 0 "
o
+ 4 0 5 6 ' 0 0 "
3 0 "
o
+ 4 0 5 5 ' 0 0 "
h m s
2 0 1 9 5 0
s
4 5
s
4 0
s
3 5
s
3 0
Fig. 2. The 2MASS Ks-band image is shown. Overplotted
are the massive stars within the diffuse emission associated
with the UC HII region IRAS 20178+4046 (see discussion
on the CM diagrams) The plus sign represents the IRAS
point source position. Contour map of the dust emission at
850 µm is also overlaid (see Sect. 4.3)
sources. The two CM diagrams show the presence of nine
early type sources with spectral types ∼ B0.5 or earlier,
out of which six sources are common to both the diagrams.
Within an error of around one subclass (except for source
# 3), the spectral type estimations derived from both the
CM diagrams are consistent. It is likely that the source # 3
(which has the least extinction among the early type stars)
is a foreground star though it is not very close to the zero
extinction curve. It is also possible that it does not belong
to the group because spatially it is farther away (see Fig. 2).
The nine early type sources identified from both the CM
diagrams are listed in Table 2.
Figure 4 shows the colour-colour (CC) diagram for the
sources in the region associated with IRAS 20178+4046.
For clarity we have classified the CC diagram into three re-
gions (e.g. Tej et al. 2006; Ojha et al. 2004a & b). The
sources in region "F" are generally considered to be ei-
ther field stars, Class III objects, or Class II objects with
small NIR excess. The "T" sources are classical T Tauri
stars (Class II objects) with large NIR excess or Herbig
AeBe stars with small NIR excess. The "P" region has
mostly protostar-like Class I objects and Herbig AeBe stars.
Considering the quoted magnitude errors in Table 2, ma-
jority of the early type sources (except source # 6 and #
3) are seen to lie in the "T" region.
In Fig. 2, we have overplotted sources listed in Table 2
on the Ks-band 2MASS image. It is interesting to see that
five of these massive stars form a cluster at the centre, of
which the position of source # 1 coincides with the IRAS
point source position. Comparing with the radio morphol-
ogy (Fig. 1 left panel), sources # 1 and # 2 are well within
the compact ionized region. Sources # 4, # 6, #8, # 9 en-
velope the UC HII region almost in a ring like structure.
Apart from this clustering, three sources (#6, #5, #7) form
a tail-type arc extending to the south. They lie within the
diffuse emission (southward lobe) seen in the Ks-band im-
age.
The presence of the massive star clustering makes the
complex interesting because the radio flux density suggests
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
5
Fig. 3. Colour-magnitude diagrams of the infrared cluster in the IRAS 20178+4046 region. The nearly vertical solid lines
represent the zero age main sequence (ZAMS) loci with 0, 15, and 30 magnitudes of visual extinction corrected for the
distance. The slanting lines show the reddening vectors for each spectral type. The magnitudes and the ZAMS loci are
all plotted in the Bessell & Brett (1988) system.
Table 2. Early type sources (∼ B0.5 and earlier) from the NIR CM diagram of IRAS 20178+4046
(hh:mm:ss.s)
Source RA (2000.0) DEC (2000.0)
(dd:mm:ss.s)
No.
+40:56:30.5
1
+40:56:35.9
2
3
+40:55:24.0
+40:56:32.6
4
5
+40:56:09.6
+40:56:25.3
6
7
+40:55:45.5
+40:56:23.7
8
+40:56:42.4
9
20:19:39.5
20:19:39.3
20:19:35.2
20:19:38.7
20:19:41.2
20:19:40.5
20:19:41.6
20:19:39.4
20:19:39.2
J
(mag)
12.491±0.034
12.532±0.030
13.380±0.022
13.400±0.040
13.930±0.026
14.118±0.046
14.405±0.037
--
--
H
(mag)
10.929±0.023
10.433±0.039
12.076±0.019
11.659±0.034
12.456±0.018
12.209±0.032
12.475±0.030
12.610±0.029
12.833±0.032
Ks
(mag)
10.016±0.027
9.150±0.029
11.580±0.015
--
11.597±0.024
11.315±0.029
11.212±0.026
11.304±0.025
11.414±0.019
Note: Magnitudes satisfying the 'read-flag' criteria of 1 -- 3 are given for the sources
a B0 -- B0.5 exciting source. It is likely that these sources
are deeply embedded. This results in absorption of a large
fraction of UV photons by the surrounding dust which re-
sults in the underestimation of the spectral type from the
radio flux density of the ionized gas. Star # 2 has spectral
type much earlier than O5 as indicated by both the CM di-
agrams. This could either be a deeply embedded pre main-
sequence (PMS) star or two or more unresolved early-type
stars. However, there is a discrepancy seen in the FIR lu-
minosity (Kurtz et al. 1994) when compared with the total
luminosity derived from the NIR sources. This could prob-
ably be due to the distance uncertainties, IR excess and
2MASS resolution limit which results in inaccuracies in the
spectral type determinations from the NIR CM diagrams.
Hence, additional infrared spectroscopic observations are
essential in determining the exact nature and spectral type
of these sources.
4.3. Dust emission in the sub-mm from JCMT-SCUBA data
We used the emission in submillimetre wavebands to study
the cold dust environment in the region around IRAS
20178+4046. The spatial distribution of cold dust emission
at 450 and 850 µm is displayed in Fig. 5. The angular res-
olutions are 8′′.8 and 14′′.0 for the 450 and 850 µm wave
bands, respectively. The 450 µm map is relatively noisy
and shows clumpy nature, whereas, the 850 µm map clearly
shows the presence of two dust cores.
The dust mass can be estimated from the following re-
lation:
Mdust = 1.88 × 10−4(cid:18) 1200
ν (cid:19)3+β
Sν(e0.048ν/Td − 1)d2 (3)
This is taken from Sandell (2000) and is a simplified ver-
sion of Eq. 6 of Hildebrand (1983). The above equation
assumes the standard Hildebrand opacities (i.e. κ1200GHz =
0.1cm2g−1). Here, Sν is the flux density at frequency ν,
6
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
2 0 "
1 0 "
o
+ 4 0 5 7 ' 0 0 "
5 0 "
4 0 "
3 0 "
2 0 "
1 0 "
o
+ 4 0 5 7 ' 0 0 "
5 0 "
4 0 "
3 0 "
2 0 "
h m s
2 0 1 9 4 1
s
4 0
s
3 9
s
3 8
s
3 7
h m s
2 0 1 9 4 2
s
4 1
s
4 0
s
3 9
s
3 8
s
3 7
Fig. 5. Contour maps showing the spatial distribution of dust emission at 450 µm (left) and 850 µm (right) for the region
around IRAS 20178+4046. The contour levels are at 35,45,55,65,75,85 and 95 % of the peak value of 4.19 and 1.72
Jy/beam at 450 and 850 µm, respectively. The FWHMs of the symmetric 2-D Gaussian beams are 8′′.8 and 14′′.0 for the
two wave bands. The plus sign and the filled triangle in each image mark the position of the IRAS point source and the
radio peak, respectively.
Td the dust temperature, β the dust emissivity index and
is taken to be 2 (Hildebrand 1983), and d the distance to
the source in kpc. We assume a typical dust temperature
of 20 K (see Tej et al. 2006; Klein et 2005). The flux den-
sities are obtained from the JCMT-SCUBA maps shown
in Fig. 5. To obtain the flux density of the entire cloud,
we integrated up to the last contour (which is at 35% of
the peak value). Using the above relation, we estimate dust
masses of ∼ 7 and 15 M⊙ from the 450 and 850 µm maps,
respectively. Assuming a gas-to-dust ratio of 100, the above
values translate to total masses of 700 and 1500 M⊙ for the
cloud from the 450 and 850 µm maps, respectively. We es-
timate the individual masses of the these two dust cores
to be ∼ 250 (northern core) and 335 M⊙ (southern core)
from the 850 µm map. The cores are defined to cover re-
gions upto ∼ 75% of the northern and southern intensity
peaks. Exploring the range of Td (20 -- 40 K) and β (1 -- 2),
we infer that the mass estimates can be lower by up to a
factor of ∼ 8.
In Fig. 6, we present the various components of the re-
gion associated with IRAS 20178+4046. The plot shows
the contour map of the dust emission at 850 µm and the
peak position of ionized gas emission at 1280 MHz over-
laid on the 3.6 µm Spitzer-IRAC4 image. As compared
to the Ks-band image, the diffuse emission seen in the
Spitzer image is far more extended. The outer envelope
spreads out to a "hat-shaped" morphology. It is interest-
ing to note that the southern core of the 850 µm map co-
incides with the radio continuum emission and the cluster
of massive stars. One MIR source (G078.4373+02.6584 --
α2000.0 = 20h19m39s.4 ; δ2000.0 = +40◦56′35′′.2), lies inside
4 The archival image presented is the Post-Basic Calibrated
Data (PBCD) downloaded using the software Leopard. This
work is based in part on observations made with the Spitzer
Space Telescope, which is operated by the Jet Propulsion
Laboratory, under NASA contract 1407.
the southern core coincident with the radio peak position.
The MSX MIR colours F21/8, F14/12, F14/8, and F21/14 of
this source fall in the zone mostly occupied by compact HII
regions (Lumsden et al. 2002). MIR emission in UIBs due
to PAH (see Appendix A) is also seen from this region. The
morphology of the UIB emission is similar to that seen in
the Spitzer image. Comparing with the FIR intensity maps
of Verma et al. (2003), it is seen that this southern core
is not spatially coincident with the FIR peaks at 150 and
210 µm but lies close to them (the core is located close to
the last contours of the maps presented in Fig. 5 of Verma
et al. 2003). It is likely that this is an evolved protocluster
or a young (partly embedded) cluster where the massive
stars develop UC HII regions and the cluster emerges out
of the parental cloud and the stars are detected in the near-
infrared. Whereas, the northern core is only detected in the
sub-mm. No radio emission (down to the level of the rms
noise in the radio maps), MIR or NIR emission is detected
for this core. As is seen clearly from Figs. 2 and 6, the north-
ern core appears as an absorption region with a notable lack
of stars. This northern core is relatively further away from
the emission peaks in the two FIR bands (150 & 210 µm) as
compared to the southern core. The mass of the two cores
are similar and satisfy the mass limit criteria of 100 M⊙
(Klein et al. 2005) for identifying earliest stages of massive
star formation. The above scenario suggests that this dense
core could be harbouring a possible pre-protocluster or an
early protocluster candidate where we are sampling the ini-
tial phases of cloud collapse in which massive stars have
possibly begun to form deeply embedded in the cluster.
Though the FIR data is not conclusive, the location of the
sub-mm dust cores with respect to the other components
most likely implies that we are seeing different evolutionary
stages of star formation in the two cores. A similar massive
star forming complex IRAS 06055+2039 was studied by Tej
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
7
et al. (2006) where the dense cloud core was at an earlier
evolutionary stage compared to the associated HII region.
4.4. Line emission from the Gas
this
star
information about
Spectroscopic signatures from the interstellar gas around
IRAS 20178+4046 have been considered to explore ad-
ditional
forming region.
Millimeter-wave observations of CS(J=2-1) by Bronfman
et al (1996), 13CO(2-1) by Wilkins et al (1989) and 13CO(1-
0) by McCutcheon et al (1991) has been translated to es-
timated gas mass of 544, 1130 and 1760 M⊙ respectively,
under reasonable assumptions. The lower value from the
CS data is expected due to a smaller beam size (39′′) used.
These masses are similar to the estimates from the thermal
emission from cold dust, 700 -- 1500 M⊙ (see Sect. 4.3). The
available infrared spectroscopic measurements for IRAS
20178+4046 includes IRAS Low Resolution Spectrometer
data (LRS; covering 8 to 22 µm) and those by Faison et
al. (1998) (covering 3 -- 5.3 & 8 -- 13 µm). The latter has
detected the [Ne II] line at 12.8 µm. We have analysed the
IRAS-LRS data to extract possible spectral signatures for
a few selected ionic fine-structure lines commonly observed
in HII regions. This resulted in clear detection of the [Ne
II] line and marginal detections of the lines due to [S III]
(18.7 µm) and [Ar III] lines (8.99 µm).
We have explored a self-consistent model of IRAS
20178+4046 region capable of explaining the observed in-
frared line emission as well as radio continuum from the
photo-ionized gas. The scheme developed by Mookerjea &
Ghosh (1999) has been used, which is based on the pho-
toionization code CLOUDY (Ferland 1996). It considers
various relevant physical and chemical processes involving
constituents of the interstellar cloud. The geometrical pa-
rameters describing the cloud were taken from Verma et
al. (2003), who have successfully modelled the observed
infrared SED and angular sizes using radiative transfer
through the dust component. Our best model corresponds
to a constant density cloud of radial dust optical depth,
τ100µm = 0.005, and a centrally embedded star cluster with
upper mass cut-off of 17.5 M⊙ (B0 type star) and the slope
of the initial mass function of -1.6. The predicted radio
emission at 1280 MHz is 51 mJy in very good agreement
with our GMRT measurement (57 mJy). The emerging in-
frared spectrum including high-contrast lines are presented
in Fig. 7 alongwith the observed SED. The inferred lumi-
nosity of the [Ne II] line from the measurements of Faison
et al. (1998) and IRAS LRS correspond to ∼ 74 & 5.3 L⊙,
respectively. This discrepancy could be due to the very dif-
ferent beams and spectral resolutions employed. The pre-
diction of our model lies in between these two values, viz.,
26.1 L⊙. The luminosities for the [S III] and [Ar III] lines
extracted from IRAS LRS are 7.7 & 0.7 L⊙ respectively,
which are lower than those predicted. However, the pre-
dicted line ratio [S III]/[Ne II] is very similar to that from
the LRS. Better measurements of these and other spectral
lines are required to resolve the above issues. Several of
the lines predicted (Fig. 7), can easily be detected by the
InfraRed Spectrograph (IRS) instrument onboard Spitzer
Space Telescope, leading to better understanding of the star
forming complex associated with IRAS 20178+4046.
Fig. 4. Colour-colour diagram of the sources in the vicinity
of the IRAS 20178+4046 region. The two solid curves repre-
sent the loci of the main sequence (thin line) and the giant
stars (thicker line) derived from Bessell & Brett (1988). The
long-dashed line is the classical T Tauri locus from Meyer
et al. (1997). The parallel dotted lines are reddening vec-
tors with the crosses placed along these lines at intervals
of five magnitudes of visual extinction. We have assumed
the interstellar reddening law of Rieke & Lebofsky (1985)
(AJ /AV = 0.282; AH /AV = 0.175 and AK /AV = 0.112).
The short-dashed line represents the locus of the Herbig
AeBe stars (Lada & Adams 1992). The plot is classified
into three regions namely "F", "T", and "P" (see text for
details). The colours and the curves shown in the figure are
all transformed to the Bessell & Brett (1988) system.
o
+ 4 0 5 9 ' 0 0 "
3 0 "
o
+ 4 0 5 8 ' 0 0 "
3 0 "
o
+ 4 0 5 7 ' 0 0 "
3 0 "
o
+ 4 0 5 6 ' 0 0 "
3 0 "
o
+ 4 0 5 5 ' 0 0 "
h m s
2 0 1 9 5 0
s
4 5
s
4 0
s
3 5
s
3 0
Fig. 6. Contour map of dust emission at 850 µm is overlaid
on the Spitzer 3.6 µm band image for the region around
IRAS 20178+4046. The filled triangle shows the position
of the radio peak, the plus sign marks the position of the
IRAS point source and the star marks the position of the
FIR peaks at 150 and 210 µm (Verma et al. 2003).
8
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
6. Conclusions
The massive star forming region associated with IRAS
20178+4046 has been studied in detail in the radio, in-
frared and sub-mm wavelengths, leading to the following
conclusions
1. High-sensitivity and high-resolution radio continuum
maps at 1280 and 610 MHz obtained from our obser-
vations using GMRT show a simple compact spherical
morphology. The total integrated emission implies an
exciting star of spectral type B0 -- B0.5.
2. The sub-mm emission from cold dust has been studied
using JCMT-SCUBA at 450 and 850 µm, from which
we estimate the total mass of the cloud to be ∼ 700
-- 1500 M⊙ which is also supported by the estimates
derived from the millimeter-wave molecular line data.
The 850 µm map shows the presence of two dense cores
of masses ∼ 250 and 335 M⊙. The southern core is most
likely an evolved protocluster or a young cluster where
the massive stars have formed the UC HII region and the
cluster emerges out of the parental cloud. The northern
core is only detected in the sub-mm and is a possible
pre-protocluster candidate.
3. The NIR colours from 2MASS data suggest the presence
of several massive stars (earlier than ∼ B0.5) within
and enveloping the UC HII region which are likely to
be deeply embedded.
4. The embedded star cluster has been characterized by
modelling the observed radio continuum and infrared
SED using a photoionization code. The upper mass cut-
off is found to be 17.5M⊙ (B0 type star) for an usual
initial mass function.
Acknowledgments
We thank the anonymous referee for providing critical com-
ments and suggestions, which have helped in improving
the scientific content of this paper. We also thank Dr.
Malcolm Walmsley for his useful suggestions. We thank
the staff at the GMRT who have made the radio obser-
vations possible. GMRT is run by the National Centre for
Radio Astrophysics of the Tata Institute of Fundamental
Research.
References
Baudry, A., Desmurs, J. F., Wilson, T. L., & Cohen, R. J. 1997, A&A,
325, 255
Bessell, M. S., & Brett, J. M. 1988, PASP, 100, 1134
Bronfman, L., Nyman, L. -A., & May, J. 1996, A&AS, 115, 81
Caswell, J. L., Murray, J. O., Roger, R. S., et al. 1975, A&A, 45, 239
Chan, S. J., Henning, T., & Schreyer, K. 1996, A&AS, 115, 285
Codella, C., Felli, M., & Natale, V. 1996, A&A, 311, 971
Codella, C., Felli, M., Natale, V., Palagi, F., Palla, F. 1994, A&A,
291, 261
Crowther, P. A., & Conti, P. S. 2003, MNRAS, 343, 143
Faison, M., Churchwell, E., Hofner, P., et al. 1998, ApJ, 500, 280
Ferland, G. J., 1996, Hazy, a brief introduction to CLOUDY, Univ.
of Kentucky, Dept. of Phys. and Astron. Internal Reports.
Garay, G., Lizano, S. 1999, PASP, 111, 1049
Ghosh, S. K., & Ojha, D. K. 2002, A&A, 388, 326
Hildebrand, R. H. 1983, QJRAS, 24, 267
Klein, R., Posselt, B., Schreyer, K., Forbich, J., & Henning, Th. 2005,
ApJS, 161, 361
Kurtz, S., Churchwell, E., & Wood, D. O. S. 1994, ApJS, 91, 659
Kurtz, S. E., Watson, A. M., Hofner, P., & Otte, B. 1999, ApJ, 514,
232
Fig. 7. The emerging spectrum (solid line) predicted by the
model for IRAS 20178+4046. The dashed line represent the
IRAS LRS from Volk & Cohen (1989). The crosses repre-
sent the photometric data taken from Verma et al. (2003)
except for the 4 MSX bands which were estimated from
respective maps.
5. General star formation scenario in IRAS
20178+4046
The morphology of the radio emission of this UC HII re-
gion at higher resolution (Kurtz et al. 1994) is cometary in
shape. As mentioned in Garay & Lizano (1999), the cham-
pagne flow and the bow shock models explain the cometary
morphologies satisfactorily though observationally it is dif-
ficult to differentiate between the two. The morphology of
the various components presented in Fig. 6 seems to suggest
a champagne flow. In this model the medium in which the
massive star is born has strong density gradients which re-
sults in the HII region moving supersonically away from the
high density region in a so-called champagne flow (Garay
& Lizano 1999). The HII region is density bounded and the
pressurized HII gas bursts out of the cloud into a fan-like or
plume-like structure. The presence of the dense dust core
and the high density to the north of the UC HII region
and the fan-like morphology seen in the Spitzer image and
the MIR image adds strength to this conjecture. It is im-
portant to note that the fan-like morphology is seen in the
other three IRAC bands also (4.5, 5.8 and 8 µm which are
not presented in this paper). Theoretical simulations in the
champagne phase by Yorke et al. (1983) show that during
this phase the resulting configuration is ionization bounded
on the high-density side and density bounded on the side of
the outward flow. The Spitzer image shows the presence of
another high density lane to the south of the fan-like diffuse
emission. This southern high density lane is spatially coin-
cident with the east-west extended emission detected in the
low resolution 8.3 GHz map of Kurtz et al. (1999). Further
high resolution observations of molecular gas velocity struc-
ture in this star forming region is crucial in understanding
the details of the environment and formation mechanism
of the massive stars. Inspite of the intrinsic temporal vari-
ability typically seen in masers (Stahler & Palla 2004) and
the sensitivity of the single dish observations carried out
for this source, the absence of the OH maser could possibly
be an indication of a more evolved HII region (Codella et
al. 1994) in support of the optically thin scenario infered
from our radio observations.
A. Tej et al.: Multiwavelength study of IRAS 20178+4046
9
5 9 '
5 8 '
5 7 '
5 6 '
5 5 '
o
+ 4 0 5 4 '
h m s
2 0 1 9 5 5
s
5 0
s
4 5
s
4 0
s
3 5
s
3 0
s
2 5
Fig. A.1. Spatial distribution of the total radiance in the
UIBs for the region around IRAS 20178+4046 as extracted
from the MSX four band images. The contour levels are at
5, 10, 20, 30, 40, 50, 60, 70, 80, 90, and 95 % of the peak
emission of 1.5 × 10−4 W m−2 Sr−1.
Lada, C. J., & Adams, F. C. 1992, ApJ, 393, 278
Lumsden, S. L., Hoare, M. G., Oudmaijer, R. D., & Richards, D. 2002,
MNRAS, 336, 621
Mathis, J. S., Mezger, P. G., Panagia, N. 1983, A&A, 128, 212
Mezger, P. G., & Henderson, A. P. 1967, ApJ, 147, 471
McCutcheon, W. H., Dewdney, P. E., Purton, C. R., & Sato, T. 1991,
AJ, 101, 1435
Meyer, M. R., Calvet, N., & Hillenbrand, L. 1997, AJ, 114, 288
Mookerjea, B. & Ghosh, S. K., 1999, J. Astrophys. Astr. 20, 1
Ojha, D. K., Tamura, M., Nakajima, Y., et al. 2004a, ApJ, 608, 797
Ojha, D. K., Tamura, M., Nakajima, Y., et al. 2004b, ApJ, 616, 1042
Panagia, N. 1973, AJ, 78, 929
Price, S. D., Egan, M. P., Carey, S. J., Mizuno, D. R., & Kuchar, T.
A. 2001, AJ, 121, 2819
Rieke, G. H., & Lebofsky, M. J. 1985, ApJ, 288, 618
Sandell, G. 2000, A&A, 358, 242
Schraml, J., & Mezger, P. G. 1969, ApJ, 156, 269
Slysh, V. I., Val'tts, I. E., Kalenskii, S. V., Voronkov, M. A., Palagi,
F., Tofani, G., & Catarzi, M. 1999, A&AS, 134,115
Stahler, S. W., Palla, F. 2004, in The Formation of Stars, WILEY-
VCH Verlag GmbH & Co. KGaA
Swarup, G., Ananthakrishnan, S., Kapahi, V. K., et al. 1991, Current
Sci., 60, 95
Szymczak, M., Hrynek, G., & Kus, A. J. 2000, A&AS, 143, 269
Tej, A., Ojha, D. K., Ghosh, S. K., Kulkarni, V. K., Verma, R. P.,
Vig, S., Prabhu, T. P. 2006, A&A, 452, 203
Verma, R. P., Ghosh, S. K., Mookerjea, B., & Rengarajan, T. N. 2003,
A&A, 398, 589
Volk, K., & Cohen, A. 1989, AJ, 98, 931
Wilking, B. A., Mundy, L. G., Blackwell, J. H., & Howe, J. E. 1989,
ApJ, 345, 257
Yorke, H. W., Tenorio-Tagle, G., Bodenheimer, P. 1983, A&A, 127,
313
Appendix A: Spatial distribution of UIBs from the
MSX data
We have used the scheme developed by Ghosh & Ojha
(2002) to extract the contribution of UIBs (due to the
PAHs) from the mid-infrared images in the four MSX
bands. The scheme models the observations with a combi-
nation of thermal emission from the 'normal' large interstel-
lar dust grains (gray body) and the UIB emission from the
gas component, under reasonable assumptions. The spatial
distribution of emission in the UIBs with an angular res-
olution of ∼ 18′′ (for the MSX survey) extracted for the
region around IRAS 20178+4046 is shown in Fig. A.1. The
morphology of the UIB emission is similar to that of the
diffuse emission seen in the Ks-band and Spitzer images
shown in Figs. 2 and 6. The integrated emission (upto 5%
of the peak value) from the UIB features in band A (viz.,
6.2, 7.7, 8.7 µm) and band C (11.3, 12.7 µm) is found to be
7.54×10−12Wm−2. In comparison, the emission in the PAH
bands (6.2, 7.7 & 11.3 µm) from the ISOCAM measure-
ments is reported to be 1.87 × 10−12Wm−2 in Verma et al.
(2003). Our larger value is not surprising considering that
it includes two additional features and also covers a larger
area. In addition, spatial distribution of warm dust com-
ponent (optical depth; τ10µm) and its temperature (TMIR)
have been generated (maps not presented here). The distri-
bution of τ10µm is compact with a peak value of 0.018, the
peak coinciding with that of the UIB emission. On the other
hand, TMIR distribution is more extended ranging between
∼ 80 -- 193 K, with the warmest region to the western edge
of extended UIB emission.
|
astro-ph/0511413 | 1 | 0511 | 2005-11-14T21:36:00 | Rubidium and lead abundances in giant stars of the globular clusters M 13 and NGC 6752 | [
"astro-ph"
] | We present measurements of the neutron-capture elements Rb and Pb in five giant stars of the globular cluster NGC 6752 and Pb measurements in four giants of the globular cluster M 13. The abundances were derived by comparing synthetic spectra with high resolution, high signal-to-noise ratio spectra obtained using HDS on the Subaru telescope and MIKE on the Magellan telescope. The program stars span the range of the O-Al abundance variation. In NGC 6752, the mean abundances are [Rb/Fe] = -0.17 +/- 0.06 (sigma = 0.14), [Rb/Zr] = -0.12 +/- 0.06 (sigma = 0.13), and [Pb/Fe] = -0.17 +/- 0.04 (sigma = 0.08). In M 13 the mean abundance is [Pb/Fe] = -0.28 +/- 0.03 (sigma = 0.06). Within the measurement uncertainties, we find no evidence for a star-to-star variation for either Rb or Pb within these clusters. None of the abundance ratios [Rb/Fe], [Rb/Zr], or [Pb/Fe] are correlated with the Al abundance. NGC 6752 may have slightly lower abundances of [Rb/Fe] and [Rb/Zr] compared to the small sample of field stars at the same metallicity. For M 13 and NGC 6752 the Pb abundances are in accord with predictions from a Galactic chemical evolution model. If metal-poor intermediate-mass asymptotic giant branch stars did produce the globular cluster abundance anomalies, then such stars do not synthesize significant quantities of Rb or Pb. Alternatively, if such stars do synthesize large amounts of Rb or Pb, then they are not responsible for the abundance anomalies seen in globular clusters. | astro-ph | astro-ph |
Rubidium and lead abundances in giant stars of the globular
clusters M 13 and NGC 67521
David Yong
Department of Physics & Astronomy, University of North Carolina, Chapel Hill, NC
27599-3255
[email protected]
Wako Aoki
National Astronomical Observatory, Mitaka, 181-8588 Tokyo, Japan
[email protected]
David L. Lambert
Department of Astronomy, University of Texas, Austin, TX 78712
[email protected]
Diane B. Paulson
NASA's Goddard Space Flight Center, Code 693.0, Greenbelt MD 20771
[email protected]
ABSTRACT
We present measurements of the neutron-capture elements Rb and Pb in
five giant stars of the globular cluster NGC 6752 and Pb measurements in four
giants of the globular cluster M 13. The abundances were derived by comparing
synthetic spectra with high resolution, high signal-to-noise ratio spectra obtained
using HDS on the Subaru telescope and MIKE on the Magellan telescope. The
program stars span the range of the O-Al abundance variation. In NGC 6752,
the mean abundances are [Rb/Fe] = −0.17 ± 0.06 (σ = 0.14), [Rb/Zr] = −0.12
± 0.06 (σ = 0.13), and [Pb/Fe] = −0.17 ± 0.04 (σ = 0.08). In M 13 the mean
abundance is [Pb/Fe] = −0.28 ± 0.03 (σ = 0.06). Within the measurement
uncertainties, we find no evidence for a star-to-star variation for either Rb or
Pb within these clusters. None of the abundance ratios [Rb/Fe], [Rb/Zr], or
[Pb/Fe] are correlated with the Al abundance. NGC 6752 may have slightly
-- 2 --
lower abundances of [Rb/Fe] and [Rb/Zr] compared to the small sample of field
stars at the same metallicity. For M 13 and NGC 6752 the Pb abundances are in
accord with predictions from a Galactic chemical evolution model. If metal-poor
intermediate-mass asymptotic giant branch stars did produce the globular cluster
abundance anomalies, then such stars do not synthesize significant quantities of
Rb or Pb. Alternatively, if such stars do synthesize large amounts of Rb or
Pb, then they are not responsible for the abundance anomalies seen in globular
clusters.
Subject headings: stars: abundances -- Galaxy: abundances -- globular clusters:
individual (M 13, NGC 6752)
1.
Introduction
Globular clusters are ideal laboratories for testing the predictions of stellar evolution the-
ory (e.g., Renzini & Fusi Pecci 1988) since the individual stars are believed to be monometal-
lic, coeval, and at the same distance. In a given globular cluster (excluding ω Cen), spec-
troscopic observations of individual stars have confirmed that members have uniform com-
positions, at least for the Fe-peak elements (e.g., see review by Gratton et al. 2004 and
references therein). However, it has been known for many years now that globular clusters
exhibit star-to-star abundance variations for the light elements C, N, O, Na, Mg, and Al
(e.g., see review by Kraft 1994). Specifically, the abundances of C and O are low when N
is high and anticorrelations are found between O and Na as well as Mg and Al. Recently,
variations in the abundance of fluorine have been discovered in giants in M 4 where the
amplitude of the dispersion exceeds that of O (Smith et al. 2005).
It is generally assumed that the light element variations arise from proton-capture re-
actions (CNO-cycle, Ne-Na chain, and Mg-Al chain), though the specific nucleosynthetic
site(s) remain elusive. One possibility for the origin of the star-to-star abundance variations
is deep-mixing and internal nucleosynthesis within the observed stars. Evidence for this
"evolutionary scenario" include C and N abundances (Suntzeff & Smith 1991) that vary
with location on the red giant branch (RGB). Extensive mixing down to very hot layers is
necessary to change the surface composition of Na, Mg, and Al. Such mixing is not predicted
1Based in part on data collected at the Subaru Telescope, which is operated by the National Astronom-
ical Observatory of Japan and on observations made with the Magellan Clay Telescope at Las Campanas
Observatory.
-- 3 --
by standard models and the proposed mechanisms include meridional circulation (Sweigart
& Mengel 1979), turbulent diffusion (Charbonnel 1995), and hydrogen-burning shell flashes
(Fujimoto et al. 1999; Aikawa et al. 2001, 2004). An alternative possibility is pollution from
intermediate-mass asymptotic giant branch stars (IM-AGBs), first suggested by Cottrell &
Da Costa (1981) to explain the Na and Al enhancements observed in CN strong stars in
NGC 6752. In IM-AGBs, hydrogen-burning at the base of the convective envelope, so-called
hot bottom burning (HBB), can produce the observed C to Al abundance patterns. Either
the ejecta from IM-AGBs pollute the proto-cluster gas from which the present cluster mem-
bers form or the ejecta are accreted by present cluster members. The strongest evidence
for this "primordial scenario" has come from observations of main-sequence stars in which
abundance variations of O, Na, Mg, and Al have been found (Gratton et al. 2001; Ram´ırez &
Cohen 2003; Cohen & Mel´endez 2005). In these unevolved stars, the internal temperatures
are too low to run the Ne-Na or Mg-Al chains which therefore precludes internal mixing as
a viable explanation for the star-to-star composition differences.
In M 4, Smith et al. (2005) found that F varied from star-to-star and that the F abun-
dance was correlated with O and anticorrelated with Na and Al. Since destruction of F
is expected to take place during HBB in IM-AGBs (Lattanzio et al. 2004), the observed
dispersion of F is in qualitative agreement with IM-AGBs being responsible for the glob-
ular cluster abundance anomalies. However, a quantitative test involving recent yields for
AGB stars combined with a standard initial mass function showed that the observed abun-
dance variations cannot be reproduced via pollution from AGB stars (Fenner et al. 2004).
Denissenkov & Herwig (2003) and Denissenkov & Weiss (2004) also find flaws in the AGB
pollution scenario based on calculated yields from AGB models. Ventura & D'Antona (2005)
caution that theoretical yields from AGB models are critically dependent upon the assumed
mass-loss rates and treatment of convection such that the predictive power of the current
AGB models is diminished. That there is still no satisfactory explanation for the star-to-
star abundance variations seen in every well studied Galactic globular cluster would suggest
that our understanding of globular cluster chemical evolution and stellar nucleosynthesis is
incomplete.
Two neutron-capture elements, rubidium and lead, may offer further clues regarding the
processes that gave rise to the star-to-star abundance variations and possibly the formation
of globular clusters. Rb has two stable isotopes, 85Rb and 87Rb. While the solar abundance
of Rb is due to 50% s-process and 50% r-process (Burris et al. 2000), the abundance of
Rb relative to nearby elements such as Sr, Y, and Zr offers an insight into the neutron
density at the site of the s-process and therefore the mass of the AGB star due to the 10.7
yr half-life of 85Kr (e.g., Tomkin & Lambert 1983, 1999; Lambert et al. 1995; Busso et al.
1999; Abia et al. 2001). Along the s-process path, Rb is preceded by Kr. The path enters
-- 4 --
at 80Kr and exits at either 85Kr or 87Kr with 85Kr providing the branching point. At low
neutron densities (Nn ≤ 1 × 108 cm−3), 85Kr β-decays to the stable isotope 85Rb. At high
neutron densities, 85Kr will capture neutrons to form 86Kr and then 87Kr which β-decays
to 87Rb (effectively stable with a half-life of 4.7 × 1010 yr). Clearly the isotopic mix of
Rb depends upon the neutron density. Unfortunately, stellar Rb isotope ratios cannot be
measured (Lambert & Luck 1976). In the presence of a steady flow along the s-process path,
the density of a nuclide satisfies the condition σiNi ≃ constant, where σi and Ni are the
cross-section and abundance of nuclide i respectively. The neutron-capture cross-sections
differ by a factor of 10 between the two Rb isotopes (σ87 = σ85/10). The 85Kr branch does
not affect the Zr abundances since the low and high neutron density s-process paths converge
at Sr. Therefore, in a high neutron density environment such as the helium intershell during
a thermal pulse in IM-AGBs, the Rb abundance may increase by a factor of 10 relative to
nearby s-process elements such as Sr, Y, and Zr. In reality, the situation is slightly more
complex since neutron capture on 84Kr leads to the ground state of 85Kr as well as a short
lived isomeric state that decays to either the 85Kr ground state or 85Rb (see Beer & Macklin
1989 for more details). IM-AGBs of solar metallicity are expected to have a high neutron
density with 22Ne(α,n)25Mg providing the neutron source. (Low-mass AGBs, whose neutron
source is 13C(α,n)16O, provide a lower neutron density.) For metal-poor or zero metallicity
IM-AGBs, Busso et al. (2001) suggest that such stars do run the s-process though the details
are model dependent. Nevertheless, the Rb abundance relative to Sr, Y, and Zr (which are
not affected by the 85Kr branch) is a potential diagnostic of the s-process site and may offer
an additional insight into the role of IM-AGBs in the chemical evolution of globular clusters.
The isotopes of Pb, along with bismuth, comprise the last stable nuclei along the s-
In low-mass AGB stars, the neutron source is provided by 13C whereas in
process path.
IM-AGBs, 22Ne provides the neutron source with the division occurring at roughly 4 M⊙.
In low-mass AGBs and IM-AGBs of low metallicity, overabundances of Pb and Bi may be
expected if the neutron supply per seed exceeds a certain value (e.g., Goriely & Mowlavi
2000; Travaglio et al. 2001; Busso et al. 2001). Goriely & Siess (2001) suggest that for AGB
stars with Z < 0.001, the available neutrons per seed nuclei is greater than the number
required to produce Pb and Bi. Travaglio et al. (2001) suggest that metal-poor IM-AGBs
play only a minor role in the production of Pb though for their 5 M⊙ model, the Pb yields
do not change between [Fe/H] = −1.3 and solar. Herwig (2004) suggest that metal-poor IM-
AGBs efficiently activate the 22Ne neutron source though quantitative s-process yields are
not presented. Busso et al. (2001) predict high yields of Pb from metal-poor IM-AGBs. The
ratio of Pb/La and Pb/Ba can be used to probe the nature of the s-process in metal-poor
AGB stars (Gallino et al. 1998; Goriely & Mowlavi 2000). Numerous observational studies
have found considerable overabundances of Pb in stars that exhibit large s- and/or r-process
-- 5 --
enhancements (e.g., Cowan et al. 1996; Sneden et al. 2000; Aoki et al. 2000, 2001, 2002; Van
Eck et al. 2001, 2003; Johnson & Bolte 2002; Lucatello et al. 2003; Sivarani et al. 2004; Ivans
et al. 2005 and references therein). If the globular cluster star-to-star abundance variations
are due to pollution from metal-poor IM-AGBs, we may expect large overabundances of Pb
and a dispersion in Pb abundances despite the absence of variations and excesses in other
s-process elements.
In this paper, we present measurements of Rb and Pb in the globular cluster NGC 6752
as well as measurements of Pb in the globular cluster M 13. While Rb has been measured
in two globular clusters, ω Cen (Smith et al. 2000) and NGC 3201 (Gonzalez & Wallerstein
1998), as far as we are aware these are the first measurements of Pb in a globular cluster. We
chose the globular clusters M 13 and NGC 6752 because they exhibit the largest amplitude
for the Al variation of all the well studied Galactic globular clusters and therefore offer
the best opportunity to find abundance variations for Rb and Pb. Previous studies of M 13
include Cohen (1978) and Peterson (1980) who found large Na variations, Shetrone (1996a,b)
who showed that the Mg isotope ratios were not constant, Cohen & Mel´endez (2005) who
discovered abundance variations in unevolved stars, as well as analyses by Kraft et al. (1992,
1997), Pilachowski et al. (1996), and Sneden et al. (2004b). Previous studies of NGC 6752
include Cottrell & Da Costa (1981) who discovered the Na and Al enhancements, Suntzeff
& Smith (1991) who found C and N to systematically vary according to evolutionary status,
Gratton et al. (2001) and Grundahl et al. (2002) who discovered O-Al variations in unevolved
stars, Yong et al. (2003a) (hereafter Y03) who found variations in Mg isotope ratios, Yong
et al. (2005) (hereafter Y05) who presented evidence for slight abundance variations of Si,
Y, Zr, and Ba, and Pasquini et al. (2005) who measured Li in main sequence stars.
2. Observations and data reduction
The list of candidates included 5 giants in NGC 6752 previously studied in Y03 and
Y05, 4 giants in M 13 previously studied by Shetrone (1996a,b), and the comparison star
HD 141531, a giant whose evolutionary status and stellar parameters are comparable to the
cluster giants. Though we were restricted to the brightest giants, the globular cluster stars
were deliberately selected to span a large range of the star-to-star abundance variations.
Table 1 contains the list of targets observed using either the Subaru or Magellan telescopes.
Observations of the M 13 giants and the comparison field star HD 141531 were obtained
with the Subaru Telescope using the High Dispersion Spectrograph (HDS; Noguchi et al.
2002) on 2004 June 1. A 0.4′′ slit was used providing a resolving power of 90,000 per 4 pixel
resolution element with wavelength coverage from 4000 A to 6700 A. For the Subaru data,
-- 6 --
one-dimensional wavelength calibrated normalized spectra were produced in the standard
way using the IRAF2 package of programs.
Observations of the NGC 6752 giants were obtained with the Magellan Telescope using
the Magellan Inamori Kyocera Echelle spectrograph (MIKE; Bernstein et al. 2003) on 2004
April 3-5. A 0.35′′ slit was used providing a resolving power of 55,000 in the red and 65,000
in the blue per 4 pixel resolution element with wavelength coverage from 3800 A to 8500 A.
While IRAF was used for most of the data reduction, extraction of the Magellan data must
account for the "tilted" slits, i.e., the lines are tilted with respect to the orders and the tilt
varies across the CCD. While this is a feature of all cross-dispersed echelle spectrographs,
for MIKE data the tilt is severe. We used the mtools3 set of tasks written by Jack Baldwin
to correct for the tilt. Failure to make this correction would result in degradation of the
spectral resolution as shown in Figure 1. (The magnitude of this effect depends upon the
aperture size applied to the order being extracted. For the stellar spectra, the decrease in
spectral resolution would be smaller than for the Th-Ar comparison spectra by roughly a
factor of 2.)
3. Analysis
3.1. Stellar parameters and the iron abundance
The first step in the analysis was to determine the stellar parameters: the effective tem-
perature (Teff), the surface gravity (log g), and the microturbulent velocity (ξt). Equivalent
widths (EWs) were measured for a set of Fe i and Fe ii lines using routines in IRAF. We
used the same set of Fe lines presented in Y03. In Figure 2, we compare the measured EWs
for Fe i and Fe ii lines for the five NGC 6752 giants analyzed in Y03. The EWs measured
in the Magellan data are in very good agreement with those measured in the VLT data.
Therefore, for the five NGC 6752 giants, we adopt the same stellar parameters used in Y03.
In Figure 3, we compare the EWs of Fe i and Fe ii lines for the comparison star HD 141531.
The EWs measured in the Subaru data are in very good agreement with those measured in
the VLT data. For HD 141531, we adopt the stellar parameters used in Y03. For the four
M 13 giants, we determined the stellar parameters using spectroscopic criteria. As in Y03
2IRAF (Image Reduction and Analysis Facility) is distributed by the National Optical Astronomy Obser-
vatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under contract
with the National Science Foundation.
3http://www.lco.cl/lco/magellan/instruments/MIKE/reductions/mtools.html
-- 7 --
and Y05, the model atmospheres were taken from the Kurucz (1993) local thermodynamic
equilibrium (LTE) stellar atmosphere grid and we used the LTE stellar line analysis program
Moog (Sneden 1973). For the effective temperature (Teff), we forced the Fe i lines to show
no trend between abundance and lower excitation potential, i.e., excitation equilibrium. To
set the surface gravity (log g), we forced the abundances from Fe i and Fe ii to be equal, i.e.,
ionization equilibrium. We adjusted the microturbulent velocity (ξt) until there was no trend
between abundance and EW. The final [Fe/H] was taken to be the mean of all Fe lines. Our
stellar parameters for the M 13 giants compare very well with those derived by Shetrone
(1996a,b), Sneden et al. (2004b), and Cohen & Mel´endez (2005).
3.2. Rubidium abundances
For the M 13 giants, the HDS spectra did not incorporate the Rb line so we were only
able to measure Rb in the NGC 6752 giants. The abundances were determined via spectrum
synthesis of the Rb i line near 7800 A (see Figures 4 and 5). Spectrum synthesis was essential
for determining accurate abundances due to hyperfine splitting and isotopic shifts as well
as blending from a stronger Si i line. While the 7800 A Rb i line is only 3-4% deep relative
to the continuum, the high quality spectra allow us to measure an abundance from this
line. Following Tomkin & Lambert (1999), the wavelengths and relative strengths for the
isotopic and hyperfine structure components were taken from Lambert & Luck (1976) and
we assumed a solar isotope ratio of 85Rb/87Rb = 3. The macroturbulent broadening was
assumed to have a Gaussian form and was estimated by fitting the profile of the nearby Ni i
line at 7798 A. We then generated synthetic spectra and varied the Si and Rb abundances
to obtain the best match to the observed spectrum. Ideally we would like to measure the Rb
isotope ratio, 85Rb/87Rb, but Lambert & Luck (1976) were unable to measure accurate ratios
in Arcturus even when using data with superior spectral resolution and signal-to-noise ratio
(S/N) due to the presence of hyperfine structure and the small isotopic shift. While our tests
confirmed that we could not measure accurate Rb isotope ratios, we verify the finding by
Tomkin & Lambert (1999) that the derived Rb abundances are not sensitive to the assumed
isotope ratio. Using the Kurucz et al. (1984) solar atlas, we measured an abundance log
ǫ(Rb) = 2.58 using a model atmosphere with Teff /log g/ξt = 5770/4.44/0.85. Our derived
solar Rb abundance is in very good agreement with the Grevesse & Sauval (1998) value, log
ǫ(Rb) = 2.60.
The subordinate Rb i line near 7947 A is weaker by a factor of 2 and is roughly 2% deep
relative to the continuum. We detect this line in all our spectra and preliminary analyses
suggest that the abundances derived from this line agree with those from the 7800 A line (see
-- 8 --
Figure 6). However, we prefer to restrict our results to the 7800 A line since the 7947 A region
is more crowded, i.e., the continuum placement strongly affects the derived Rb abundances
from the 7947 A line. Furthermore, unidentified blends, atmospheric absorption, and fringing
are more prevalent in this spectral region. (CN lines lie in this region may be absent in these
metal-poor globular cluster stars.)
In the subsequent sections, we compare our globular cluster Rb abundances with field
and cluster stars with [Fe/H] > −2.0 analyzed by other investigators. We now attempt to
place the various Rb abundance measurements onto a common scale. While the isotope ratio
of Rb cannot be measured in Arcturus, the elemental abundance ratio is well known. Using
the Hinkle et al. (2000) Arcturus atlas, we measured an abundance [Rb/H] = −0.55 using
a model atmosphere with Teff/log g/ξt = 4300/1.5/1.55 obtained using the spectroscopic
criteria described in Section 3.1. The adopted Rb gf value was identical to that used by
Tomkin & Lambert (1999) and our derived Rb abundance is in very good agreement with
their measured value, [Rb/H] = −0.58. Since Arcturus is common to both studies and our
derived abundances are essentially identical, we therefore make no adjustment to the Tomkin
& Lambert Rb abundances. Gratton & Sneden (1994) use the same gf value, though the
relative strengths of the hyperfine components differ slightly. We do not adjust their Rb
abundances. Abia et al. (2001) adopt a Rb gf value that differs from ours, so we adjust
their Rb abundances by +0.08 dex. It is not clear what Rb gf was used by Gonzalez &
Wallerstein (1998). Fortunately, Arcturus was also part of their sample. They derived an
abundance [Rb/H] = −0.45 and so we adjust their Rb abundances by −0.1 dex. Smith et al.
(2000) find [Rb/H] = −0.52 and so we do not adjust their Rb abundances.
3.3. Lead abundances
The Pb abundances were determined via spectrum synthesis of the Pb i line near
4058 A (see Figures 7 and 8). Abundances from the Pb line near 3683 A could not be
determined due to the lack of flux in the blue for these cool giants. While the region cen-
tered near 4058 A is crowded with molecular lines of CH as well as atomic lines from Mg,
Ti, Mn, Fe, and Co, our syntheses provide a very good fit to the region demonstrating that
reliable Pb abundances can be extracted. The macroturbulent broadening was estimated by
fitting the profiles of the nearby lines. We adopted the same gf value used by Aoki et al.
(2000, 2001, 2002). Following Aoki et al. (2002), our synthesis accounted for the hyperfine
and isotopic splitting as well as the isotopic shifts. (The stable isotopes are 204Pb, 206Pb,
207Pb, and 208Pb.) We again assumed a solar isotope ratio for Pb though as with Rb, our
tests confirmed that the derived elemental Pb abundance was not sensitive to this choice.
-- 9 --
For the solar Pb abundance, we adopted log ǫ(Pb) = 1.95 from Grevesse & Sauval (1998).
In the subsequent sections, we compare the globular cluster Pb abundances with those
obtained in field stars with [Fe/H] > −2.0 by other investigators. However, we further
restrict the comparison by avoiding stars with known s-process enhancements leaving only
a handful of stars from two studies, Sneden et al. (1998) and Travaglio et al. (2001). Again
we attempt to put the Pb abundance measurements onto a common scale. Our gf value
is identical to that used by Sneden et al. (1998) so we do not adjust their Pb abundances.
Travaglio et al. (2001) used a different Pb line (3683 A) and they did not list the adopted
gf value. Since there are no stars common to both analyses, we do not make an adjustment
to their Pb abundances and caution that there may be a systematic offset.
3.4. Additional elements
We also measured abundances for Al, Si, Y, Zr, La, and Eu in the M 13 giants and the
comparison star HD 141531 using the same lines presented in Y05. These measurements
were performed to ensure that the abundances would be on the same scale as the NGC 6752
giants studied in Y05. Zr was chosen because we compare the Rb and Zr abundances to
look for a large ratio [Rb/Zr] as well as a detectable dispersion, i.e., the hallmark of a high
neutron-density environment and the possible signature of pollution from IM-AGBs. Al, Si,
and Y were also chosen because Y05 found evidence for correlations between Al and Si, Al
and Y, and Al and Zr in NGC 6752. While our sample size in M 13 is small, it would be
interesting to see if similar correlations are present. La and Eu were measured since these
neutron-capture elements offer an insight into the ratio of s-process to r-process material.
Furthermore, the ratio [Pb/La] has been used to test predictions from AGB models.
In
Table 2 we present our measured elemental abundances for Al, Si, Rb, Y, Zr, La, Eu, and
Pb in the program stars. The adopted solar abundances for Al, Si, Y, Zr, La, and Eu were
6.47, 7.55, 2.24, 2.60, 1.13, and 0.52 respectively.
We attempt once more to put the abundance measurements onto a common scale by
considering the gf values used by the various studies to which we compare our abundances.
The element we focus upon is Zr (in order to compare [Rb/Zr] between the samples). We
shift all the Zr abundances onto the Smith et al. (2000) scale in order to compare with their
theoretical predictions for [Rb/Zr] from low and intermediate-mass AGBs (their Figure 14).
(Our Rb abundances were already on the Smith scale.) Abundance measurements for Zr
are complicated by the fact that the laboratory Zr gf values from Bi´emont et al. (1981) are
smaller than the solar gf -values by 0.41 dex (Tomkin & Lambert 1999). We must therefore
take care and account for both the adopted solar abundance and the gf value. Smith et al.
-- 10 --
(2000) adopt the Bi´emont et al. (1981) gf -values and a solar abundance log ǫ⊙(Zr) = 2.90.
We used the Bi´emont et al. (1981) gf -values and a solar abundance log ǫ⊙(Zr) = 2.60
(Grevesse & Sauval 1998). Therefore we adjust our Zr abundances by −0.30 dex to ensure
that we are on the same scale as Smith et al. (2000). Similarly, we adjust the Zr abundances
of Gratton & Sneden (1994) and Abia et al. (2001) by −0.30 dex since they adopt the same
gf values and a very similar solar abundance used in our analysis. Gonzalez & Wallerstein
(1998) adopt the Grevesse & Sauval (1998) solar abundance and a different gf value so we
adjust their Zr abundances by +0.06 dex. Tomkin & Lambert (1999) adopt the Grevesse
& Sauval (1998) solar abundance and a different gf value so we adjust their Zr abundances
by +0.11 dex. These adjustments are substantial. When we compare the ratio [Rb/Zr], we
also consider how the comparison would fare had we not made these abundance corrections.
To assess the validity of these adjustments, we measured the Zr abundance for Arcturus and
found [Zr/H] = −0.66. Smith et al. measured [Zr/H] = −0.96, Gonzalez & Wallerstein found
−1.18, and Tomkin & Lambert found −1.00. Therefore, applying the abundance corrections
based on the gf values and solar abundances ensures that the Zr abundances are on the
Smith et al. (2000) scale, e.g., for Arcturus we find −0.96 (this study), −1.12 (Gonzalez &
Wallerstein), −0.89 (Tomkin & Lambert), and −0.96 (Smith et al.). Interestingly, if we had
used the same solar abundance as Smith et al. (2000), our [Zr/Fe] abundances in Y05 would
have been closer to the [Y/Fe] values and for M 13 we would have found [Y/Fe] ≃ [Zr/Fe].
In Figure 9, we plot [Zr/Fe] versus [Fe/H] with the abundances shifted to the Smith et al.
(2000) scale. At the metallicity of NGC 6752 and M 13, the field and cluster stars have
similar ratios of [Zr/Fe]. The comparison field star HD 141531 has [Zr/Fe] almost identical
to the globular clusters.
As in Y03, we estimate the internal errors in the stellar parameters to be Teff ± 50K,
log g ± 0.2, and ξt ± 0.2. In Table 3, we show the abundance dependences upon the model
parameters.
4. Discussion
4.1. Rubidium
While our mean Rb abundance for NGC 6752 is [Rb/Fe] = −0.17 ± 0.06 (σ = 0.14),
the abundances appear to concentrate around two distinct values. There are two stars with
[Rb/Fe] ≃ −0.02 and three stars with [Rb/Fe] ≃ −0.25. The two stars with the higher Rb
abundances do not have the highest Al abundances and the three stars with the lower Rb
abundances are not exclusively the stars with the lowest Al abundances. Given the weakness
of the Rb line, the uncertainties in the derived Rb abundances (see Table 3), and the small
-- 11 --
sample size, it is unlikely that the Rb abundances show a dispersion in NGC 6752. Nor do
we find evidence for a correlation between [Al/Fe] and [Rb/Fe], though we recognize that our
sample size (5 stars) is much more limited than in Y03 and Y05 (38 stars). Unfortunately,
observations of M 13 and the comparison field star HD 141531 did not incorporate the Rb
line.
In Figure 10, we compare our Rb abundances with those measured in dwarfs and giants
in the disk and halo (Gratton & Sneden 1994 and Tomkin & Lambert 1999), globular cluster
giants (Gonzalez & Wallerstein 1998 and Smith et al. 2000), and carbon-rich AGB stars
(Abia et al. 2001). (While we retain their stars in the plots, we note that the Abia sample
contains very different objects that are difficult to analyze compared to the dwarfs and
giants considered in the other studies.) Recall that we have made small adjustments to the
Rb abundances in an attempt to place them onto a common scale. At the metallicity of
NGC 6752 ([Fe/H] = −1.6), our two stars in NGC 6752 with the highest [Rb/Fe] ratios have
abundances compatible with the lower envelope of the Tomkin & Lambert (1999) sample.
The two NGC 6752 stars also exhibit very similar abundances [Rb/Fe] to the Gonzalez &
Wallerstein (1998) and Smith et al. (2000) globular cluster giants. Our three stars with the
lower [Rb/Fe] ratios appear unusual compared to the Tomkin & Lambert sample. Only 1
star in the Abia et al. (2001) sample has [Fe/H] < −1.0 and it is interesting that it has an
abundance ratio [Rb/Fe] similar to those measured in NGC 6752. In general, Rb is not a
well studied element and the comparison data are limited.
For NGC 6752, we find a mean abundance [Rb/Zr] = −0.12 ± 0.06 (σ = 0.13). (This
abundance ratio has been shifted to the Smith et al. (2000) scale.) For the five NGC 6752
giants, the ratio [Rb/Zr] appears to show a dispersion. We suspect that this is attributable
to measurement uncertainties (primarily for Rb) rather than reflecting a real star-to-star
scatter. We do not find a correlation between [Al/Fe] and [Rb/Zr]. In Figure 11, we compare
the abundance ratio [Rb/Zr] between NGC 6752 and various field and cluster stars. Note
that in this Figure we have shifted all abundances onto the Smith et al. (2000) scale since
we will utilize their theoretical predictions from low and intermediate-mass AGBs. At the
metallicity of NGC 6752, we find that the two stars in NGC 6752 with the highest values
of [Rb/Fe] also have the highest values of [Rb/Zr]. These two stars have similar [Rb/Zr]
ratios to the Gratton & Sneden (1994) and Tomkin & Lambert (1999) samples at the same
metallicity. While the ω Cen giants have abundance ratios [Rb/Zr] slightly lower than NGC
6752, this time the NGC 3201 giants appear to have much higher ratios of [Rb/Zr]. Note
that the [Rb/Zr] ratios in NGC 3201 appear similar to the highest values seen in the Tomkin
& Lambert (1999) sample. Unfortunately, the only star in the Abia et al. (2001) sample
with [Fe/H] < −1.0 does not have a Zr measurement. However, it does have [Rb/Sr] = −0.5
and [Rb/Y] = −0.6. If we assume for this star [Rb/Zr] = <[Rb/Sr],[Rb/Y]> = −0.55, then
-- 12 --
the abundance is much lower than NGC 6752.
For elements heavier than Si, globular clusters and field stars tend to have very similar
abundance ratios [X/Fe] at a given [Fe/H] (Gratton et al. 2004; Sneden et al. 2004a). Al-
though the scatter is large and the sample sizes are limited, it would appear that cluster stars
probably have similar, or perhaps slightly lower abundance ratios of [Rb/Fe] and [Rb/Zr]
compared to field stars at a given [Fe/H].
Recall that we made substantial adjustments to the Zr abundance. While consideration
of the Arcturus Zr abundances would appear to validate this adjustment, we briefly consider
how the comparison of [Rb/Zr] would have fared if these corrections were not applied. In
this case, the ratio [Rb/Zr] would decrease by roughly 0.3 dex for NGC 6752 as well as
for the Gratton & Sneden (1994) and Abia et al. (2001) samples. The Tomkin & Lambert
(1999) and Gonzalez & Wallerstein (1998) samples would increase by roughly 0.15 dex. NGC
6752 would therefore have unusually low ratios [Rb/Zr] compared to field stars at the same
metallicity. Similarly, the ω Cen compositions would be unusually low though comparable
to NGC 6752. The NGC 3201 giants would then have very high [Rb/Zr] ratios compared to
other globular clusters and field stars at the same metallicity. NGC 3201 is peculiar since
it has a retrograde orbit and may have been a captured cluster (van den Bergh 1993). The
capture hypothesis could not be demonstrated by Gonzalez & Wallerstein (1998) who found
no unusual abundance ratios.
Smith et al. (2000) compare [Rb/Zr] in ω Cen with predictions from AGB models
with various initial masses and initial metallicities (their Figure 14). Their Figure clearly
shows how the ratio [Rb/Zr] can vary by nearly a factor of 10 depending on whether a
low-mass (1.5M⊙) or high-mass (5M⊙) AGB model is synthesizing the s-process elements.
As anticipated from the arguments given in Section 1, high-mass AGB models produce high
[Rb/Zr] while low-mass AGB models produce low [Rb/Zr]. The magnitude of the difference
in the predicted [Rb/Zr] between low- and high-mass AGB models does not significantly
change as the metallicity decreases from [Fe/H] = −0.5 to [Fe/H] = −2.0. Comparing the
observed abundances with the model predictions in Smith et al. reveals that low-mass AGB
stars (1-3 M⊙) are responsible for the synthesis of the s-process elements in ω Cen. Inspection
of Figure 14 in Smith et al. (2000) shows that at the metallicity of NGC 6752, [Fe/H] =
−1.6, our measured ratio [Rb/Zr] = −0.12 is compatible with the s-process elements being
synthesized in low-mass AGB stars though the assumed mass of the 13C pocket is critical.
Predictions assuming a standard treatment for the 13C pocket or the 13C pocket increased
by a factor of 2 both suggest AGB stars with < 3 M⊙ are responsible for the [Rb/Zr] ratios
seen in NGC 6752. When the 13C pocket is diminished by a factor of 3, the AGB stars with
masses > 3 M⊙ may explain the observed [Rb/Zr]. When we return our Zr abundances to
-- 13 --
the original scale, [Rb/Zr] = −0.42, the ratio in NGC 6752 is only compatible with low-mass
AGB stars. We note that the highest values of [Rb/Zr] seen in the Gratton & Sneden (1994),
Gonzalez & Wallerstein (1998), and Tomkin & Lambert (1999) samples all greatly exceed
the 5 M⊙ AGB model predictions. Such a discrepancy serves as a useful reminder of the
unfortunate reality that the detailed yields of s-process elements from AGB stars may be
very model dependent (Busso et al. 2001; Ventura & D'Antona 2005).
4.2. Lead
In NGC 6752, the mean Pb abundance is [Pb/Fe] = −0.17 ± 0.04 (σ = 0.08) and
in M 13 the mean abundance is [Pb/Fe] = −0.28 ± 0.03 (σ = 0.06). Given the fairly
large measurement uncertainty for Pb (see Table 3), neither NGC 6752 nor M 13 show any
evidence for a dispersion in Pb abundances, though our sample sizes for both clusters are
small. Furthermore, the [Pb/Fe] ratios are very similar for these two clusters. As with Rb,
there is no evidence that the ratio [Pb/Fe] is correlated with [Al/Fe]. We note that one star,
NGC 6752 PD1, has lower ratios of both [Rb/Fe] and [Pb/Fe] relative to other giants in this
cluster. This subtle composition difference probably arises from uncertainties in the stellar
parameters rather than representing a genuine difference. The comparison field giant HD
141531 has a ratio [Pb/Fe] essentially identical to the globular cluster giants.
In Figure 12, we compare our Pb abundances with values measured by Sneden et al.
(1998) and Travaglio et al. (2001). While Pb has been measured in numerous stars with
large s-process enhancements, it has been largely neglected in normal field stars presumably
due to the difficulty of the measurement. For HD 126238, the Pb abundance measured by
Sneden et al. (1998) is very similar to the globular cluster giants and HD 141531. The Pb
abundances measured by Travaglio et al. (2001) in field stars are larger than those measured
in the globular clusters. The star with [Pb/Fe] = 0.6 is a CH star with excess C and Ba and
should not be considered a normal field star. Aside from the CH star, there are three stars
with upper limits and another three Pb detections. Recall that there are no stars common
to both studies and that the Pb gf value was not published. Travaglio et al. (2001) suggest
that some of the Pb detections may be uncertain and therefore, the offset between the Pb
abundances may be due to measurement errors and/or the adopted gf value.
Travaglio et al. (2001) not only measured Pb abundances in a handful of stars, but they
calculated the Galactic chemical evolution of Pb from a detailed model. In their Figure 4,
they plot the expected run of [Pb/Fe] versus [Fe/H] for the halo, thick disk, and thin disk.
Since their prediction integrates over all AGB masses, it would be useful to learn how the
predicted curve would differ (if at all) if the calculation was performed using low-mass or
-- 14 --
high-mass AGB models exclusively as Smith et al. (2000) have done. At [Fe/H] = −1.6, the
Travaglio et al. (2001) model predicts an abundance ratio [Pb/Fe] ≃ −0.1. This prediction
is in very good agreement with the values measured in M 13, NGC 6752, HD 126238, and
HD 141531. This agreement may be regarded as evidence that globular cluster stars have
virtually identical Pb abundances as normal field stars.
In normal field stars, Pb has been less studied than Rb. Clearly, it would be of great
interest to have additional Pb measurements in field and cluster stars. In Figures 7 and 8,
our syntheses indicate that for cool giants in the metallicity regime −2.0 < [Fe/H] < −1.0,
reliable Pb abundances can be measured even in stars that do not have large Pb or s-process
enhancements.
The ratio [Pb/La] may offer further clues regarding the nature of the s-process in the
AGB stars. Van Eck et al. (2003) found some stars with ratios of [Pb/La] > +1.5, in
agreement with predictions from metal-poor AGB models (Gallino et al. 1998; Goriely &
Mowlavi 2000). However, Aoki et al. (2002) and Van Eck et al. (2003) also found a large
spread in the ratio [Pb/Ba]. In some stars, the ratio [Pb/Ba] was sub-solar. Our mean ratio
[Pb/La] for M 13 is −0.36 ± 0.05 (σ = 0.10). For NGC 6572, our mean ratio [Pb/La] =
−0.23 ± 0.04 (σ = 0.09). In both clusters, the mean ratios are similar and we note that they
are both sub-solar and comparable to the lowest ratios seen in the Van Eck et al. (2003)
sample. Curiously the subsample in Van Eck et al. (2003) with [Pb/La] < 0 had extreme
enhancements for [Pb/Fe] and [La/Fe]. The comparison field star HD 141531, has [Pb/La]
= −0.20 which is similar to the value seen in the globular clusters.
4.3. Additional elements
While our sample in M 13 consists of only 4 stars, they span the extremities of the Al
variation. As in Y05, we again find that the most Al-rich stars may also exhibit slightly
higher Si abundances than the most Al-poor stars. Further measurements of Al and Si in
a large sample of stars in M 13 would be of great interest to verify whether the correlation
between Al and Si seen in NGC 6752 (Y05) is also present in M 13. The correlations between
Al and Y as well as Al and Zr found in NGC 6752 do not appear to be present in the small
M 13 sample. Cohen & Mel´endez (2005) measured abundances in 25 stars in M 13 from the
main sequence turn-off to the tip of the RGB. They were unable to measure Al in most stars.
When we consider their derived abundances, there appears to be an anticorrelation between
O and Si as well as O and Y. Though the anticorrelation is driven primarily by the one star
with unusually low [O/Fe], such trends are intriguing and warrant further investigation.
-- 15 --
As noted in previous investigations of these clusters, the ratio of s-process to r-process
material, [La/Eu], is sub-solar but greater than the scaled solar pure r-process value. For
NGC 6752 and M 13, the observed ratios of [La/Eu] show that AGBs have contributed to
their chemical evolution. The ratio of La/Eu in HD 141531 again confirms that it is a normal
field star.
4.4. Consequences for the IM-AGB pollution scenario
In Y03, we measured Mg isotope ratios in bright giants in NGC 6752. We found that
the ratio varied from star-to-star. Specifically, 24Mg was anticorrelated with Al, 26Mg was
correlated with Al, and 25Mg was not correlated with Al. As previously seen by Shetrone
(1996b) in M 13, these isotope ratios reveal that the Al enhancements result from proton
capture on the abundant 24Mg. Proton capture on 24Mg within the Mg-Al chain is predicted
to only occur in AGB stars of the highest mass at their maximum luminosity (Karakas &
Lattanzio 2003). So we suggested that the abundance variations were due to differing degrees
of pollution by IM-AGBs, an idea originally proposed by Cottrell & Da Costa (1981). These
IM-AGBs must have the same iron abundance as the present generation of cluster stars
otherwise there would also be a star-to-star abundance variation of Fe. (The same argument
applies to whatever stars are believed to be the source of the pollutants.)
The Mg isotope ratios presented in Y03 offered further clues to globular cluster chemical
evolution. At one extreme of the abundance variation are cluster stars with O, Na, Mg, and
Al compositions in accord with field stars at the same metallicity. We called such stars
"normal" in anticipation that proton capture nucleosynthesis can produce O-poor, Na-rich,
Mg-poor, and Al-rich material. At the other extreme of the abundance variations are the
stars with high Na, high Al, low O, and low Mg. We referred to these stars as "polluted".
The pollution may have occurred via either the evolutionary or primordial scenario.
In
"normal" stars, we found ratios 25Mg/24Mg and 26Mg/24Mg that exceeded field stars at the
same metallicity (Yong et al. 2003b). Of equal importance was the fact that these isotope
ratios greatly exceeded predictions from metal-poor supernovae. We therefore suggested that
these unusually high isotope ratios could be explained if a previous generation of IM-AGBs
of the highest mass polluted the natal cloud from which the cluster formed. The ejecta from
this previous generation must have been thoroughly mixed before the present generation
of stars began to form. This previous generation of IM-AGBs are probably responsible for
much of the Na, Al, and N as well as 25Mg and 26Mg.
Our working hypothesis is that IM-AGBs played two crucial roles in globular cluster
chemical evolution. Firstly, a prior generation of very metal-poor IM-AGBs are required to
-- 16 --
produce the high 25Mg/24Mg and 26Mg/24Mg seen in "normal" stars. Secondly, a generation
of IM-AGBs with the same Fe abundance as the present cluster members pollutes the cluster
environment. Differing degrees of pollution of natal clouds then produce the star-to-star
abundance variations. The dispersion in the F abundances and the correlation between F
and O in M 4 (Smith et al. 2005) appear to confirm the role of IM-AGBs in producing
the abundance variations. However, not all the abundance patterns observed in globular
clusters can be matched by the current theoretical yields from IM-AGBs (Denissenkov &
Herwig 2003; Denissenkov & Weiss 2004) nor can the abundance patterns be reproduced by
chemical evolution models (Fenner et al. 2004).
From a qualitative viewpoint, metal-poor IM-AGBs may produce s-process elements
via the 22Ne neutron source. If activated, the 22Ne neutron source produces large amounts
of Rb/Zr due to a critical branching point at 85Kr as described earlier. Theoretical models
by Busso et al. (2001) suggest that metal-poor IM-AGBs do run the s-process though the
specific yields depend on the details.
Similarly, a qualitative assessment suggests that metal-poor IM-AGBs will produce Bi
In this case, Bi and Pb
and Pb if the neutrons per seed nuclei exceed a certain value.
may show large enhancements with other s-process elements showing only modest overabun-
dances. Again, theoretical models can be found in which metal-poor IM-AGBs do produce
lead (e.g., Goriely & Siess 2001 and Busso et al. 2001).
In M 13 and NGC 6752, we did not find high ratios of [Rb/Fe], [Rb/Zr], or [Pb/Fe]
compared with field stars at the same metallicity. If metal-poor IM-AGBs are responsible
for the globular cluster star-to-star abundance variations, then our measurements strongly
suggest that such stars do not synthesize significant quantities of Rb or Pb. Also, if metal-
poor IM-AGBs are responsible for the large abundances of 25Mg and 26Mg in "normal"
cluster stars, then they do not synthesize Rb or Pb. Alternatively, if metal-poor IM-AGBs
do synthesize significant quantities of Rb and Pb, then they cannot be responsible for the
abundance anomalies seen in globular clusters.
Of course the possibility remains that the predicted yields of Rb and Pb from IM-
AGBs are unreliable and/or model dependent (e.g., Ventura & D'Antona 2005). It has been
suggested that metal-poor IM-AGBs will not produce any s-process elements. As the mass of
the AGB star increases, the size of the He intershell region decreases as does the duration of
the thermal pulse (Lattanzio et al. 2004). Detailed predictions of the yields from metal-poor
IM-AGBs by independent groups are required. Our observed Rb and Pb abundances may
serve to constrain these models.
Finally, we note that Rb and Zr are synthesized in IM-AGBs as well as via the weak
-- 17 --
s-process, i.e., He core burning in massive stars. However, Travaglio et al. (2004) suggest
that the weak s-process does not contribute to Zr but a lighter element primary process in
massive stars may be responsible for up to 18% of the solar abundance of Zr. Chieffi &
Limongi (2004) present detailed yields from massive stars for a range of metallicities and
masses. At Z = 0, massive stars are predicted to produce low ratios, [Rb/Zr] = −0.5 to −2.4.
If IM-AGBs have contributed to the chemical evolution of NGC 6752, they would increase
the ratios of [Rb/Zr] above those produced by the supernovae. If IM-AGBs have not played a
role in the chemical evolution, perhaps we can use [Rb/Zr] to probe the mass and metallicity
range of the previous generation of massive stars. For Z = 10−4, the predicted yields are
independent of mass with [Rb/Zr] = −0.23. However, we note that for other metallicities,
there is a mass-metallicity degeneracy for the [Rb/Zr] yields that limits their use in probing
the previous generation of supernovae.
5. Concluding remarks
We show for the first time the uniformity of the neutron-capture elements Rb and Pb in
NGC 6752 and M13, the two globular clusters that exhibit the largest dispersion for Al. We
also find the ratio [Rb/Zr] to be constant. None of the abundance ratios [Rb/Fe], [Rb/Zr],
or [Pb/Fe] are correlated with [Al/Fe] and the Rb and Pb abundances show sub-solar ratios
[X/Fe]. If metal-poor IM-AGBs produce large amounts of Pb and Rb as well as high ratios
of [Rb/Zr], then such stars are not responsible for the abundance variations, a conclusion
already suggested by Denissenkov & Herwig (2003), Denissenkov & Weiss (2004), and Fenner
et al. (2004). If metal-poor IM-AGBs are responsible for the abundance variations, then they
cannot produce overabundances of Rb or Pb.
For elements heavier than Al, previous studies have shown that field and cluster stars
generally have the same abundance ratios [X/Fe] at a given [Fe/H]. While our sample size is
small and the data for comparison field stars are limited, the two clusters we have studied
have Rb abundance ratios [Rb/Fe] and [Rb/Zr] in reasonable agreement with the general
field population (the clusters may have slightly lower ratios). The Pb abundance ratio
[Pb/Fe] in globular clusters is in very good agreement with the limited sample of field stars.
At the metallicity of M 13 and NGC 6752, their Pb abundances are well matched by the
predictions from the chemical evolution model by Travaglio et al. (2001). In order to further
our understanding of stellar nucleosynthesis and the chemical evolution of field and cluster
stars, additional measurements of Rb and Pb in normal stars and globular clusters are
welcomed as are further theoretical efforts to calculate the yields from metal-poor IM-AGBs.
-- 18 --
This research has made use of the SIMBAD database, operated at CDS, Strasbourg,
France and NASA's Astrophysics Data System. DY thanks John Lattanzio and Roberto
Gallino for helpful discussions, Chris Sneden for providing a linelist for the Pb region, and
Bruce Carney for a thorough review of a draft of this paper. This research was performed
while DBP held a National Research Council Research Associateship Award at NASA's God-
dard Space Flight Center. DLL acknowledges support from the Robert A. Welch Foundation
of Houston, Texas. This research was supported in part by NASA through the American
Astronomical Society's Small Research Grant Program.
REFERENCES
Abia, C., Busso, M., Gallino, R., Dom´ınguez, I., Straniero, O., & Isern, J. 2001, ApJ, 559,
1117
Aikawa, M., Fujimoto, M. Y., & Kato, K. 2001, ApJ, 560, 937
-- . 2004, ApJ, 608, 983
Aoki, W., Norris, J. E., Ryan, S. G., Beers, T. C., & Ando, H. 2000, ApJ, 536, L97
Aoki, W., Ryan, S. G., Norris, J. E., Beers, T. C., Ando, H., Iwamoto, N., Kajino, T.,
Mathews, G. J., & Fujimoto, M. Y. 2001, ApJ, 561, 346
Aoki, W., Ryan, S. G., Norris, J. E., Beers, T. C., Ando, H., & Tsangarides, S. 2002, ApJ,
580, 1149
Beer, H. & Macklin, R. L. 1989, ApJ, 339, 962
Bernstein, R., Shectman, S. A., Gunnels, S. M., Mochnacki, S., & Athey, A. E. 2003, in
Instrument Design and Performance for Optical/Infrared Ground-based Telescopes.
Edited by Iye, Masanori; Moorwood, Alan F. M. Proceedings of the SPIE, Volume
4841, pp. 1694-1704 (2003)., 1694 -- 1704
Bi´emont, E., Grevesse, N., Hannaford, P., & Lowe, R. M. 1981, ApJ, 248, 867
Burris, D. L., Pilachowski, C. A., Armandroff, T. E., Sneden, C., Cowan, J. J., & Roe, H.
2000, ApJ, 544, 302
Busso, M., Gallino, R., Lambert, D. L., Travaglio, C., & Smith, V. V. 2001, ApJ, 557, 802
Busso, M., Gallino, R., & Wasserburg, G. J. 1999, ARA&A, 37, 239
-- 19 --
Charbonnel, C. 1995, ApJ, 453, L41
Chieffi, A. & Limongi, M. 2004, ApJ, 608, 405
Cohen, J. G. 1978, ApJ, 223, 487
Cohen, J. G. & Mel´endez, J. 2005, AJ, 129, 303
Cottrell, P. L. & Da Costa, G. S. 1981, ApJ, 245, L79
Cowan, J. J., Sneden, C., Truran, J. W., & Burris, D. L. 1996, ApJ, 460, L115+
Denissenkov, P. A. & Herwig, F. 2003, ApJ, 590, L99
Denissenkov, P. A. & Weiss, A. 2004, ApJ, 603, 119
Fenner, Y., Campbell, S., Karakas, A. I., Lattanzio, J. C., & Gibson, B. K. 2004, MNRAS,
353, 789
Fujimoto, M. Y., Aikawa, M., & Kato, K. 1999, ApJ, 519, 733
Gallino, R., Arlandini, C., Busso, M., Lugaro, M., Travaglio, C., Straniero, O., Chieffi, A.,
& Limongi, M. 1998, ApJ, 497, 388
Gonzalez, G. & Wallerstein, G. 1998, AJ, 116, 765
Goriely, S. & Mowlavi, N. 2000, A&A, 362, 599
Goriely, S. & Siess, L. 2001, A&A, 378, L25
Gratton, R., Sneden, C., & Carretta, E. 2004, ARA&A, 42, 385
Gratton, R. G., Bonifacio, P., Bragaglia, A., Carretta, E., Castellani, V., Centurion, M.,
Chieffi, A., Claudi, R., Clementini, G., D'Antona, F., Desidera, S., Fran¸cois, P., Grun-
dahl, F., Lucatello, S., Molaro, P., Pasquini, L., Sneden, C., Spite, F., & Straniero,
O. 2001, A&A, 369, 87
Gratton, R. G. & Sneden, C. 1994, A&A, 287, 927
Grevesse, N. & Sauval, A. J. 1998, Space Science Reviews, 85, 161
Grundahl, F., Briley, M., Nissen, P. E., & Feltzing, S. 2002, A&A, 385, L14
Herwig, F. 2004, ApJ, 605, 425
-- 20 --
Hinkle, K., Wallace, L., Valenti, J., & Harmer, D. 2000, Visible and Near Infrared Atlas of
the Arcturus Spectrum 3727-9300 A (Visible and Near Infrared Atlas of the Arcturus
Spectrum 3727-9300 A ed. Kenneth Hinkle, Lloyd Wallace, Jeff Valenti, and Dianne
Harmer. (San Francisco: ASP) ISBN: 1-58381-037-4, 2000.)
Ivans, I. I., Sneden, C., Gallino, R., Cowan, J. J., & Preston, G. W. 2005, ApJ, 627, L145
Johnson, J. A. & Bolte, M. 2002, ApJ, 579, L87
Karakas, A. I. & Lattanzio, J. C. 2003, Publ. Astron. Soc. Australia, 20, 279
Kraft, R. P. 1994, PASP, 106, 553
Kraft, R. P., Sneden, C., Langer, G. E., & Prosser, C. F. 1992, AJ, 104, 645
Kraft, R. P., Sneden, C., Smith, G. H., Shetrone, M. D., Langer, G. E., & Pilachowski, C. A.
1997, AJ, 113, 279
Kurucz, R. 1993, ATLAS9 Stellar Atmosphere Programs and 2 km/s grid. Kurucz CD-ROM
No. 13. Cambridge, Mass.: Smithsonian Astrophysical Observatory, 1993., 13
Kurucz, R. L., Furenlid, I., & Brault, J. 1984, Solar flux atlas from 296 to 1300 NM (National
Solar Observatory Atlas, Sunspot, New Mexico: National Solar Observatory, 1984)
Lambert, D. L. & Luck, R. E. 1976, The Observatory, 96, 100
Lambert, D. L., Smith, V. V., Busso, M., Gallino, R., & Straniero, O. 1995, ApJ, 450, 302
Lattanzio, J., Karakas, A., Campbell, S., Elliott, L., & Chieffi, A. 2004, Memorie della
Societa Astronomica Italiana, 75, 322
Lucatello, S., Gratton, R., Cohen, J. G., Beers, T. C., Christlieb, N., Carretta, E., &
Ram´ırez, S. 2003, AJ, 125, 875
Noguchi, K., Aoki, W., Kawanomoto, S., Ando, H., Honda, S., Izumiura, H., Kambe, E.,
Okita, K., Sadakane, K., Sato, B., Tajitsu, A., Takada-Hidai, T., Tanaka, W., Watan-
abe, E., & Yoshida, M. 2002, PASJ, 54, 855
Pasquini, L., Bonifacio, P., Molaro, P., Francois, P., Spite, F., Gratton, R. G., Carretta, E.,
& Wolff, B. 2005, A&A in press (astro-ph/0506651)
Peterson, R. C. 1980, ApJ, 237, L87
Pilachowski, C. A., Sneden, C., Kraft, R. P., & Langer, G. E. 1996, AJ, 112, 545
-- 21 --
Ram´ırez, S. V. & Cohen, J. G. 2003, AJ, 125, 224
Renzini, A. & Fusi Pecci, F. 1988, ARA&A, 26, 199
Shetrone, M. D. 1996a, AJ, 112, 1517
-- . 1996b, AJ, 112, 2639
Sivarani, T., Bonifacio, P., Molaro, P., Cayrel, R., Spite, M., Spite, F., Plez, B., Andersen,
J., Barbuy, B., Beers, T. C., Depagne, E., Hill, V., Fran¸cois, P., Nordstrom, B., &
Primas, F. 2004, A&A, 413, 1073
Smith, V. V., Cunha, K., Ivans, I. I., Lattanzio, J. C., & Hinkle, K. H. 2005, ApJ in press
(astro-ph/0506763)
Smith, V. V., Suntzeff, N. B., Cunha, K., Gallino, R., Busso, M., Lambert, D. L., & Straniero,
O. 2000, AJ, 119, 1239
Sneden, C. 1973, ApJ, 184, 839
Sneden, C., Cowan, J. J., Burris, D. L., & Truran, J. W. 1998, ApJ, 496, 235
Sneden, C., Cowan, J. J., Ivans, I. I., Fuller, G. M., Burles, S., Beers, T. C., & Lawler, J. E.
2000, ApJ, 533, L139
Sneden, C., Ivans, I. I., & Fulbright, J. P. 2004a, in Origin and Evolution of the Elements,
172
Sneden, C., Kraft, R. P., Guhathakurta, P., Peterson, R. C., & Fulbright, J. P. 2004b, AJ,
127, 2162
Suntzeff, N. B. & Smith, V. V. 1991, ApJ, 381, 160
Sweigart, A. V. & Mengel, J. G. 1979, ApJ, 229, 624
Tomkin, J. & Lambert, D. L. 1983, ApJ, 273, 722
-- . 1999, ApJ, 523, 234
Travaglio, C., Gallino, R., Arnone, E., Cowan, J., Jordan, F., & Sneden, C. 2004, ApJ, 601,
864
Travaglio, C., Gallino, R., Busso, M., & Gratton, R. 2001, ApJ, 549, 346
van den Bergh, S. 1993, ApJ, 411, 178
-- 22 --
Van Eck, S., Goriely, S., Jorissen, A., & Plez, B. 2001, Nature, 412, 793
-- . 2003, A&A, 404, 291
Ventura, P. & D'Antona, F. 2005, A&A in press (astro-ph/0505221)
Yong, D., Grundahl, F., Lambert, D. L., Nissen, P. E., & Shetrone, M. D. 2003a, A&A, 402,
985
Yong, D., Grundahl, F., Nissen, P. E., Jensen, H. R., & Lambert, D. L. 2005, A&A, 438, 875
Yong, D., Lambert, D. L., & Ivans, I. I. 2003b, ApJ, 599, 1357
This preprint was prepared with the AAS LATEX macros v5.2.
-- 23 --
Fig. 1. -- Profiles of extracted ThAr lines from the same exposure taken with Magellan-MIKE
scaled to the same peak intensity. The solid black line shows the ThAr lines extracted without
taking into account the tilted slits. The dashed red line shows the ThAr lines extracted using
the mtools package. Note how the FWHM decreases when the correction is performed.
-- 24 --
Fig. 2. -- Comparison of the equivalent widths of Fe I and Fe II lines between this study
(Magellan data) and Y03 (VLT data) for the sample of NGC 6752 giants.
-- 25 --
Fig. 3. -- Comparison of the equivalent widths of Fe I and Fe II lines between this study
(Subaru data) and Y03 (VLT data) for HD 141531.
-- 26 --
Fig. 4. -- Spectrum of NGC 6752 B702 near the 7800 A Rb feature. Synthetic spectra with
differing Rb abundances are shown.
-- 27 --
Fig. 5. -- Same as Figure 4 but for star NGC 6752 B1630.
-- 28 --
Fig. 6. -- Spectrum of NGC 6752 B708 near the 7947 A Rb feature. Synthetic spectra with
differing Rb abundances are shown. The lower panel is identical to the upper panel except
for the y-range.
-- 29 --
Fig. 7. -- Spectrum of NGC 6752 B708 near the 4058 A Pb feature. Synthetic spectra with
differing Pb abundances are shown.
-- 30 --
Fig. 8. -- Same as Figure 7 but for star M 13 L973.
-- 31 --
Fig. 9. -- [Zr/Fe] versus [Fe/H]. Closed black circles show our measurements for NGC 6752,
the open black circles show M 13, the open black square shows HD 141531, the closed green
triangles are from Gratton & Sneden (1994), the green asterisks represent data from Gonzalez
& Wallerstein (1998), open blue triangles are from Tomkin & Lambert (1999), red plus signs
represent data from Smith et al. (2000), and filled red squares show data from Abia et al.
(2001). A representative error bar is shown and the Zr abundances have been shifted onto
the Smith et al. scale (see text for details).
-- 32 --
Fig. 10. -- [Rb/Fe] versus [Fe/H]. Closed black circles show our measurements for NGC 6752,
the closed green triangles are from Gratton & Sneden (1994), the green asterisks represent
data from Gonzalez & Wallerstein (1998), open blue triangles are from Tomkin & Lambert
(1999), red plus signs represent data from Smith et al. (2000), and filled red squares show
data from Abia et al. (2001). A representative error bar is shown.
-- 33 --
Fig. 11. -- [Rb/Zr] versus [Fe/H]. The symbols are the same as in Figure 10. Note that we
shifted all abundances onto the Smith et al. (2000) scale (see text for details).
-- 34 --
Fig. 12. -- [Pb/Fe] versus [Fe/H]. Closed black circles, open blue triangles, and the closed
red square show our measurements for NGC 6752, M 13, and the comparison field star HD
141531. The closed green triangles and green arrows (upper limits) are from Travaglio et al.
(2001) while the purple asterisk shows HD 126238 from Sneden et al. (1998). A representative
error bar is shown.
-- 35 --
Table 1. Exposure times and stellar parameters.
Star
Telescope
Exposure
Time (min)
S/Na
S/Na
4050A 7800A
Teff
K
log g
ξt
[Fe/H]
km s−1
60
50
60
64
20
50
26
27
21
10
39
37
39
36
42
55
47
44
46
85
. . .
. . .
. . .
. . .
197
402
197
247
240
. . .
3900
3950
3950
3920
4050
4050
3928
3900
3894
4273
0.00
0.20
0.30
0.30
0.50
0.25
0.26
0.24
0.33
0.80
2.25
2.25
2.25
2.35
2.10
2.20
2.70
2.70
2.50
1.90
−1.56
−1.63
−1.59
−1.61
−1.58
−1.63
−1.62
−1.60
−1.59
−1.72
Subaru
M 13 L598
Subaru
M 13 L629
M 13 L70b
Subaru
M 13 L973c
Subaru
NGC 6752 B702 Magellan
NGC 6752 B708 Magellan
NGC 6752 PD1
Magellan
NGC 6752 B1630 Magellan
NGC 6752 B3589 Magellan
HD 141531
Subaru
aS/N values are per pixel.
bAlternative name II-67.
cAlternative name I-48.
Table 2. Elemental abundances.
Star
[Al/Fe]
[Si/Fe]
[Rb/Fe]
[Y/Fe]
[Zr/Fe]a
[Zr/Fe]b
[La/Fe]
[Eu/Fe]
[Pb/Fe]
M 13 L598
M 13 L629
M 13 L70
M 13 L973
NGC 6752 B702
NGC 6752 B708
NGC 6752 PD1
NGC 6752 B1630
NGC 6752 B3589
HD 141531
0.24
0.74
1.27
1.28
1.04
1.23
1.08
0.82
0.77
0.01
0.23
0.31
0.33
0.35
0.43
0.35
0.38
0.41
0.43
0.23
. . .
. . .
. . .
. . .
−0.02
−0.22
−0.25
−0.03
−0.32
. . .
−0.10
−0.05
−0.12
−0.05
0.04
0.04
0.07
0.04
0.10
−0.13
0.27
0.25
0.24
0.17
0.32
0.30
0.23
0.19
0.20
0.14
−0.03
−0.05
−0.06
−0.13
0.02
0.00
−0.07
−0.11
−0.10
−0.16
0.00
0.06
0.14
0.12
0.09
0.03
0.07
0.05
0.08
0.00
0.40
0.44
0.43
0.52
0.28
0.28
0.32
0.34
0.35
0.33
−0.30
−0.20
−0.27
−0.35
−0.10
−0.15
−0.30
−0.10
−0.18
−0.20
aZr abundances using the gf values and solar abundance assumed in Y05.
bZr abundances when shifted onto the Smith et al. (2000) scale.
-- 36 --
Table 3. Abundance dependences on model parameters for NGC 6752 B1630.
Species
Teff + 50K log g + 0.2cgs
ξt + 0.2km s−1
[Fe/H]
[Al/Fe]
[Si/Fe]
[Rb/Fe]
[Y/Fe]
[Zr/Fe]
[La/Fe]
[Eu/Fe]
[Pb/Fe]
0.02
0.05
−0.01
0.09
0.01
0.13
−0.02
−0.04
0.12
0.01
−0.01
0.03
0.03
0.06
0.02
0.05
0.05
−0.05
−0.03
−0.01
−0.01
−0.01
−0.04
−0.01
−0.05
−0.04
−0.02
|
astro-ph/9605204 | 1 | 9605 | 1996-06-03T01:16:58 | Internal Kinematics of Distant Field Galaxies: I. Emission Line Widths for a Complete Sample of Faint Blue Galaxies at <z>=0.25 | [
"astro-ph"
] | We present measurements of the OII(3727) emission line width for a complete sample of 24 blue field galaxies (21.25<B<22, B-R<1.2) at <z>=0.25, obtained with the AUTOFIB fibre spectrograph on the Anglo-Australian Telescope. Most emission lines are spectrally resolved, yet all have dispersions sigma<100km/s. Five of the 24 sample members have OII doublet line flux ratios which imply gas densities in excess of 100 cm^-3. The line emission in these galaxies may be dominated by an active nucleus and the galaxies have been eliminated from the subsequent analysis. The remaining 19 linewidths are too large by a factor of two (7sigma significance) to be attributed to turbulent motions within an individual star forming region, and therefore most likely reflect the orbital motion of ionized gas in the galaxy. We use Fabry--Perot observations of nearby galaxies to construct simulated datasets that mimic our observational setup at z=0.25; these allow us to compute the expected distribution of (observable) linewidths sigma_v for a galaxy of a given ``true'' (optical) rotation speed v_c. These simulations include the effects of random viewing angles, clumpy line emission, finite fibre aperture, and internal dust extinction on the emission line profile. We assume a linewidth--luminosity--colour relation: ln[ v_c(M_B,B-R) ] = ln[v_c(-19,1)] - eta*(M_B+10) + zeta*[(B-R)-1] and determine the range of parameters consistent with our data. We find a mean rotation speed of v_c(-19,1)=66+-8km/s (68% confidence limits) for the distant galaxies with M_B=-19 and B-R=1, with a magnitude dependence for v_c of eta=0.07+-0.08, and a colour dependence of zeta =0.28+-0.25. Through comparison with several local samples we show that this value of v_c(-19,1) is significantly lower than the optical rotation speed of present-day galaxies with the same absolute magnitude | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000–000 (2012)
Internal Kinematics of Distant Field Galaxies:
I. Emission Linewidths for a Complete Sample of
Faint Blue Galaxies at hz i ∼ 0.25
Hans-Walter Rix1 , Puragra Guhathakurta2 , Matthew Colless3 & Kristine Ing2
1 Max-Planck-Institut fur Astrophysik, 85740 Garching, Germany
and Steward Observatory, Univ. of Arizona, Tucson, Az 85721, USA
2 UCO/Lick Observatory, University of California, Santa Cruz, California 95064, USA
3 Mount-Stromlo and Siding Springs Observatory, The Australian National University, Canberra, ACT 2611, Australia
22 July 2012
ABSTRACT
We present measurements of the [Oii] emission line width for a com-
plete sample of 24 blue field galaxies (21.25 < B < 22, B − R < 1.2)
at hz i ∼ 0.25, obtained with the AUTOFIB fibre spectrograph on the
Anglo-Australian Telescope (AAT). Most emission lines are spectrally
resolved, yet all have dispersions σv < 100 km s−1 . Five of the 24 sam-
ple members have [Oii] doublet line flux ratios which imply gas den-
sities in excess of 100 cm−3 . The line emission in these galaxies may
be dominated by an active nucleus and the galaxies have been elimi-
nated from the subsequent analysis. The remaining 19 linewidths are
too large by a factor of two (7σ significance) to be attributed to tur-
bulent motions within an individual star forming region, and therefore
most likely reflect the orbital motion of ionized gas in the galaxy. We
use Fabry–Perot observations of nearby galaxies to construct simulated
datasets that mimic our observational setup at z ∼ 0.25; these allow us
to compute the expected distribution of (observable) linewidths σv for
2
Rix, Guhathakurta, Col less, Ing
a galaxy of a given “true” (optical) rotation speed vc . These simulations
include the effects of random viewing angles, clumpy line emission, fi-
nite fibre aperture, and internal dust extinction on the emission line
profile. We assume a linewidth–luminosity–colour relation:
loghvc (cid:0)MB , B −R(cid:1)i = log [vc (−19, 1)] − η (cid:0)MB + 19(cid:1) + ζ [(B −R)− 1]
and determine the range of parameters consistent with our data. We
find a mean rotation speed of vc (−19, 1) = 66 ± 8 km s−1 (68% confi-
dence limits) for the distant galaxies with MB = −19 and B − R = 1,
with a magnitude dependence for vc of η = 0.07 ± 0.08, and a colour
dependence of ζ = 0.28 ± 0.25. Through comparison with several lo-
cal samples we show that this value of vc (−19, 1) is significantly lower
than the optical rotation speed of present-day galaxies with the same
absolute magnitude and rest frame colour (≈ 105 km s−1 ). The most
straightforward interpretation is that the distant blue, sub-L∗ galaxies
are about 1.5 mag brighter (and ≥ 0.8 mag brighter at 99% confidence)
than local galaxies of the same linewidth and colour.
Key words: galaxies: evolution, kinematics, fundamental parameters – cosmology: observations
1 INTRODUCTION
In the last decade, large ground-based telescopes, HST and sensitive detectors have
made it possible to study galaxy properties over cosmological distances in order to
obtain direct constraints on galaxy evolution. One of the most surprising results of
these studies has been the high number counts of “blue” field galaxies at magnitudes
fainter than B ∼ 20 (Kron 1980; Tyson 1988). Even though these counts suggest
significant evolutionary effects, the shape of the measured redshift distribution of
blue-selected galaxies (B ≤ 23) is consistent with no-evolution models (Broadhurst
et al. 1988; Colless et al. 1990, hereafter referred to as CETH; Lilly et al. 1991; Colless
et al. 1993b). Further, deep imaging surveys in the red or near infrared bandpasses
Linewidths of Distant Field Galaxies
3
yield galaxy counts which do not indicate strong evolutionary effects (Cowie et al.
1993; Gardner et al. 1993).
These puzzling observational results have prompted a large number of models to
explain the observations. These models fall into two broad classes in their approach.
First, there are ad hoc models, which assume some evolution of galaxy properties
(number density, luminosity, etc.) and test which of these scenarios are consistent with
the observations. The evolutionary effects in these models are not directly related to
the physics of galaxy formation. The no-evolution hypothesis, mild and strong lumi-
nosity and spectral evolution models (Tinsley 1972; Guiderdoni & Rocca-Volmerange
1990; Gronwall & Koo 1995), the differential luminosity evolution model (Broadhurst
et al. 1988), the rapidly fading/disappearing dwarf galaxy picture (Babul & Rees
1991), and number density evolution models (cf. the merging scenario of Broadhurst
et al. 1992) fall into the first category. A second class of evolutionary models tries to
incorporate a substantial amount of independent information about the hierarchical
formation of structure in the Universe (cf. Kauffmann et al. 1994; Cole et al. 1994).
There the number density of collapsed, virialized structures is derived as a function
of epoch from a given cosmogonic scenario (e.g., a cold dark matter cosmogony), and
combined with prescriptions for the merging and star-formation history of galaxies.
The luminosity and density evolution of galaxies are coupled in this latter class of
models.
Even though these models differ widely in many aspects, they share a common
feature: the bulk of the evolutionary changes are attributed to intrinsically faint
(L < 0.5 L∗ ), blue galaxies. This feature can explain the paucity of galaxies with
z >∼ 1 in redshift surveys of B -selected galaxies, and may explain why evolutionary
effects appear to be stronger in observations at shorter wavelengths. The observed
number counts and redshift distribution constrain the overall evolution of the galaxy
luminosity function, but do not provide a unique prescription for the evolutionary
history of individual galaxies. Therefore it is still a wide open question to which
4
Rix, Guhathakurta, Col less, Ing
extent the evolution of the luminosity function is caused by changes in the luminosity
of individual galaxies or by changes in number density of visible galaxies.
The key to understanding the evolution of individual galaxies lies in the identifi-
cation of the local counterparts of distant galaxies. In this paper we try to establish
a link between distant and local galaxies by comparing the relation between the
spatially integrated Doppler linewidth (LW) of the ionized gas to the continuum lu-
minosity in the two populations. This observational test effectively seeks to determine
how the luminosity of galaxies of a given LW or mass scale—i.e., the specific lumi-
nosity of galaxies—changes from moderate redshifts to the present day, and thereby
addresses the question of “luminosity evolution” directly.
Various workers have shown that it is feasible to obtain line-width measure-
ments and rotation curves for galaxies at cosmologically significant distances (Vogt et
al.1993, Colless 1994, Koo et al. 1995; Rix, Colless and Guhathakurta, 1995; Guzman
et al. 1996; Vogt et al. 1996). The are three important new elements in this paper:
(1) a statistically well defined sample, (2) detailed modeling of the observations, and
(3) a careful comparison with properly chosen local galaxy samples.
We have carried out an experiment to determine the global LW of the ionized
gas—using the [Oii] doublet—in a complete, magnitude-limited sample of blue field
galaxies at hzi ∼ 0.25. Using a fibre spectrograph, we have performed the optical
analog of a single beam Hi LW measurement. The target galaxies have the magni-
tudes and colours of local, small, star-forming spiral galaxies (similar to the Large
Magellanic Cloud), and we compare the measured LWs of these target galaxies with
the LWs of probable local counterparts, after correcting for several biases inherent in
such measurements.
The paper is organized as follows: Sec. 2 contains a description of the target sam-
ple, observations, and data reduction, and a discussion of how the LWs are measured
from the data. In Sec. 3, we interpret the measured LW is terms of the characteristic
rotation speed of the galaxy, and study the correlation of the galaxies’ rotation speed
with their absolute luminosity and colour. In Sec. 4, we compare the rotation speeds
Linewidths of Distant Field Galaxies
5
of distant galaxies to those of local counterparts in order to quantify the evolution of
field galaxies. The main points of this paper are summarized in Sec. 5.
A Hubble constant of H0 = 70 km s−1 Mpc−1 and flat space-time (q0 = 0.5) are
assumed throughout.
2 SAMPLE SELECTION, OBSERVATIONS, AND DATA
REDUCTION
2.1 The Target Sample
The target galaxies were chosen at random from the photometric database used by
CETH for the LDSS-1 survey, on the basis of the following criteria: (a) apparent blue
magnitudes in the range 21.25 ≤ bJ ≤ 22, and (b) colours bluer than bJ − rF ≤ 1.2.
(For the sake of convenience, we use the symbols B and R in the rest of the paper
when referring to magnitudes in the bJ and rF bands, as defined in CETH.)
In
addition to the magnitude and colour criteria used to define the target sample, the
final sample of LWs includes only galaxies with moderate to strong [Oii] emission
(equivalent widths >∼ 10 A) within the observed redshift range 0.16 < z < 0.37, which
corresponds to the spectral window for [Oii] in our setup. In Sec. 2.4, we demonstrate
that for most galaxies that satisfy criteria (a) and (b) the requirement of detectable
[Oii] emission does not impose any additional restriction.
Our colour cut corresponds to the rest frame colours of Scd galaxies at a redshift
of 0.25 (cf. Fig. 12 in CETH), and has been designed to yield a high fraction of galaxies
with detectable emission lines. The photometric database from which the galaxies for
CETH’s LDSS-1 redshift survey were drawn (Jones et al. 1991) contains 764 galaxies
in our chosen apparent magnitude range, 21.25 < B < 22, and 463 (= 61% ± 3%) of
these are in our chosen colour range, B − R < 1.2. The colour distribution of nearly
2000 galaxies with 21 < B < 22 in Kron’s (1980) photometric sample indicates that
>∼ 50% of the galaxies within this apparent magnitude range survive our colour cut.
The median galaxy in our sample (B = 21.7 at z = 0.25) has an absolute blue
6
Rix, Guhathakurta, Col less, Ing
B = −20.6 for H0 = 70 km s−1 Mpc−1
magnitude of MB ∼ −19.0, or 0.25 L∗
B [L∗
(Efstathiou et al. 1988)]. The absolute magnitudes of the galaxies listed in Table 1
were calculated from their apparent B magnitudes, accounting for their luminosity
distance and a K-correction for Scd galaxies (cf. Fig. 13.7 in Peebles 1993).
2.2 Observations
The high-resolution spectra presented in this paper were obtained with the AUT-
OFIB multifibre spectrograph at the Anglo-Australian Telescope (AAT) on 1993
October 12, with seeing of ≈ 1 .′′7. A 1200 lines mm−1 grating was used to cover
the spectral region from 4300 A to 5110 A at a dispersion of 0.79 A pixel−1 . The
wavelength coverage was chosen to detect redshifted [Oii] (3727 A) emission line dou-
blets in galaxies with 0.16 < z < 0.37, a redshift range centred on the peak of the
redshift distribution of B = 21.25–22 galaxies to maximize the [Oii] detection rate.
The instrumental resolution is σ = 0.82 A (FWHM = 1.9 A), which corresponds to
σ = 56 km s−1 and 48 km s−1 at the blue and red ends of the spectral range, re-
spectively. The doublet nature of the [Oii] emission line, with a doublet separation
of ∆λ = 2.75 A, is resolved at the instrumental resolution. This makes it necessary
to tailor the data reduction process (Sec. 2.5), but has the advantage of allowing an
unambiguous redshift determination based on only one emission line.
The spectra were recorded on a 1024 × 1024 Tektronix CCD. The AUTOFIB
spectrograph contains 60 usable fibres, each 2 .′′ 1 in diameter, which can be positioned
over a 40′ diameter circular field. Engineering tests indicate that the fibres can be
positioned with an accuracy of 0 .′′3 rms. The overall throughput of the system is a
little under 1%. Of the 60 fibres, 49 were positioned on target galaxies in the FBG1
field. The approximate coordinates of the field center are: α1950 = 00h54m51s ; δ1950 =
−27◦51′ .
Five of the remaining fibres were positioned at random ‘blank’ sky locations in
order to measure (and to subsequently subtract) the night sky spectrum, while six
were pointed at nearby bright stars that served as astrometric references. The total
Linewidths of Distant Field Galaxies
7
exposure time on the FBG1 field was 7 hr, split into individual 1 hr exposures. The
1 hr target exposures were interspersed with short (2 min) calibration exposures of
Cu–Ar and Fe–Ar arc lamps. The seeing was about 1 .′′ 7 (FWHM) as measured off the
astrometric reference stars which were observed through fibre bundles. Bias frames
and short flat field exposures of the twilight sky and of a tungsten lamp (white light
source) were also obtained.
2.3 Data Reduction
The fibre spectra were reduced using standard procedures in the IRAF packages ccd-
proc and dohydra. The individual 1 hr CCD frames were flat fielded, the wavelength
scale in the FBG1 spectra calibrated using arc lamp spectra, and the data combined
after rejection of cosmic ray events. A one-dimensional, optimally weighted, mean
spectrum was extracted for each fibre from the combined CCD data, using the flat
field image to define the response profile of each fibre and the locus of its spectrum on
the CCD array. The spectra are dominated by night sky continuum emission outside
the [Oii] lines, and the lines occupy a very small fraction (< 1%) of the full spectral
range; the continuum flux is undetectable for most galaxies. The median of all spec-
tra (galaxy and blank sky) was used to define an accurate template of the night sky
spectrum and it was subtracted from each galaxy’s spectrum. The sky-subtracted
spectra of all 24 galaxies in which [Oii] emission is detected are shown in Figure 1.
2.4 Detection Limits for [OII] Emission
2.4.1 Detection Algorithm
We can devise an ob jective detection criterion for the emission lines, because galaxies
with clearly detected [Oii] will not have any other emission line in the observed
spectral range; the portion of the spectrum outside the [Oii] line therefore provides
an empirical estimate of the effective noise in the data, including Poisson noise, sky
subtraction error, residual cosmic rays, and flat fielding error.
8
Rix, Guhathakurta, Col less, Ing
We convolve the spectrum of each galaxy with a sequence of template [Oii] line
profiles, each centered on the k -th pixel (with 20 ≤ k ≤ 1004, the portion of the
CCD free from edge artifacts). The templates are double Gaussians of equal height
and instrumental linewidth (σinstr = 0.82 A), broadened by various amounts to ac-
count for the intrinsic galaxy LW. The correlation amplitude C (k), resulting from
the convolution of these templates with the data, has a strong maximum Cmax at
the redshift of any obvious [Oii] emission line. We calculate the rms variation, ∆C ,
of C (k) from the rest of the spectrum and determine the ratio of the second highest
maximum of C (k) to ∆C . Since this second maximum must be due to noise, we can
determine how high to set a threshold to avoid such spurious detections. Applying
this statistic to all spectra with “clear” detections, i.e., with Cmax/∆C ≥ 10, indi-
cates that a threshold Cthresh/∆C = 4.5 avoids any false detection. Using a narrow
instr + (40 km s−1 )2 ]1/2 , we searched all 49 target galaxy
template of width, σ = [σ2
spectra, detecting 24 galaxies with Cmax > Cthresh (4.5 ∆C ). A histogram of detection
significances is given in Figure 3 and the observed properties of these galaxies are
listed in Table 1⋆ .
As narrow emission lines have a higher contrast against the galaxy’s continuum
and sky background than broad lines, they are easier to detect. To check how our
detection limit depends on LW, we have repeated the above procedure, broadening
the template by 0 km s−1 < σv ≤ 300 km s−1 . The convolution with broader templates
resulted in no additional detections over those listed in Table 1. Further, all but one
of the existing [Oii] lines could have been detected even if their LWs had been as
large as σv ≤ 200 km s−1(for a given total line flux). We conclude that our detection
efficiency is approximately uniform for σv ≤ 200 km s−1 . The fact that all of the
24 galaxies detected have LWs σv < 100 km s−1 is not the result of a detection bias.
⋆ Note that for convenience we use units of both A and km s−1 for the dispersion σ; the ratio of these two quantities
is λ/c for the instrumental dispersion, σinstr , and λ0 /c for the intrinsic rest frame dispersion of the [Oii] lines, σv .
Linewidths of Distant Field Galaxies
9
2.4.2 Completeness
There are two independent arguments that we have detected most, possibly all, of
the target galaxies (21.25 < B < 22, B − R < 1.2) within the searched redshift range
(0.16 < z < 0.37).
(1) The strength of the emission line in most detected ob jects, as measured by the
value of Cmax , is considerably higher than the detection threshold Cthresh = 4.5∆C .
The weakest detection is 5.4 ∆C , and the median detection amplitude is 11 ∆C ,
as shown in Figure 3. If most of the true [Oii] line fluxes for this sample of blue
galaxies were below our detection limit and if we just saw the tip of the iceberg,
we would have expected to find the ma jority of detected galaxies very close to the
limit Cmax ≈ Cthresh .
(2) Complete redshift surveys (CETH; Colless et al. 1993a) show that about 45%±10%
of all field galaxies in the magnitude range of our target sample (21.25 < B < 22)
have redshifts between 0.16 and 0.37. Our detection rate, 24 out of 49, is consistent
with this. Given the Poisson error in the number of galaxies (20%) and large
scale clustering of faint galaxies (e.g., the redshift “spikes” seen in the survey by
Broadhurst et al. 1988), we conclude that we have detected at least 75% of the
target galaxies within our surveyed redshift range, and possibly all of them.
Hence we believe that we are measuring LWs for a well-defined, representative sample
of distant field galaxies.
2.5 Measuring the [OII] Linewidths
In measuring linewidths for the galaxies in our sample, we must account for several
factors:
• We do not know the shape of the line profile a priori. This is because the shape
of the rotation curve is unknown, because the emission from ionized gas is lumpy
and often asymmetric with respect to the dynamical center of the host galaxy, and
10
Rix, Guhathakurta, Col less, Ing
because not all of it comes from the flat part of the rotation curve (cf. Schommer
et al. 1993; also Sec. 3.1).
• The [Oii] (3727A) line is a doublet whose line flux ratio R[OII] depends on the
electron density in the line emitting gas. This ratio must be determined from the
data.
• The rest frame separation of the [Oii] doublet [(∆λ)0 = 2.75 A] is not much larger
than the instrumental resolution and/or the typical kinematic line broadening
observed in our galaxy sample.
We cope with these complications by choosing a specific model for the line profile of
the doublet:
[λ − λz ǫ ]2
[λ − λz ]2
instr ] o#,
Imod (λ) = I0 "exp n−
instr ] o + R[OII] exp n−
z ǫ + σ2
z + σ2
2[σ2
2[σ2
with λz ≡ (1 + z)λ0 , λz ǫ ≡ (1 + z)(1 + ǫ)λ0 , σz ≡ (1 + z)σv , σz ǫ ≡ (1 + z)(1 + ǫ)σv , and
λ0 = 3726.05 A. This form incorporates the known doublet separation expressed as a
fractional difference, ǫ = (∆λ)0/λ0 = 7.3805 × 10−4 , and ensures identical rest frame
velocity dispersions, σv (in λ units), for the two doublet components. It approximates
the instrumental line profile by a Gaussian whose width, σinstr = 0.82 A, has been
(1)
determined from the widths of Cu–Ar and Fe–Ar arc lamp lines.
By choosing the functional form given in Eq. (1), we define the galaxy’s LW as the
σ of the best fitting Gaussian to the observed emission line profile. The parameters to
be fit are the line amplitude I0 , the [Oii] doublet flux ratio R[OII] , the redshift z , and
the rest frame intrinsic dispersion of the emission line σv (σv [A] = σv [km s−1 ]λ0/c).
Further, a linear continuum baseline is fitted simultaneously over the 80 pixels neigh-
bouring the emission line.
The best fit parameters (I0 , R[OII] , z , σv ) are determined by minimizing χ2 ,
!2
χ2 = Xpixel k Imod (λ) − Ik
∆Ik
directly in pixel space (cf. Rix & White 1992). The quantities Ik and ∆Ik are the
(2)
,
observed intensity and noise levels, respectively, at the k -th pixel with wavelength λ.
Linewidths of Distant Field Galaxies
11
The “1σ” error bars on the best fit values of the individual parameters are determined
by finding the region of parameter space for which χ2 − χ2
min = 1. The upper and
lower error bars are calculated separately. The two error estimates are comparable for
most parameters and for most ob jects, except for low values of σv (see upper panel
of Fig. 2) and we tabulate the geometric mean of the two error estimates, ∆obs
v
.
The best fit line parameters for all 24 galaxies are given in Table 1. The corre-
sponding fits are superimposed onto the data in Figure 1; Figure 2 shows two of the
fits in detail.
The line luminosities, L[OII] , and equivalent widths, EW, listed in Table 1 are
not based on flux calibrations. Instead, they have been calculated from the exposure
time, assuming an overall instrumental efficiency of 0.9%. Non-photometric conditions
during the observations and inaccuracies in the fibre throughput calibration result
in a factor of two uncertainty in the determination of L[OII] and EW. Note that
for positive definite quantities such as the dispersion, the “best fit” value is not an
unbiased estimator of the true value when the error is comparable to the measured
value (cf. Wardle & Kronenberg 1974). In this regime, the probability distribution
can be approximated by a mean of σv ≡ qσ2
v )2 and a dispersion of ∆obs
v, fit − (∆obs
v
where ∆obs
is the geometric mean of the upper and lower errors derived from the χ2
v
fit. The best fit values of the velocity dispersion, σv, iit, are presented in Table 1 and
,
in Figure 4; In the analysis of Sec. 3, however, we use the corrected values. Note that
errorbars, ∆obs
v
, listed in Table 1, Figure 4 and Equation 5 are the geometric mean
of the upper and lower errorbars.
The main observational results based on these fits are displayed in Figure 4,
which shows several properties of the 24 galaxies with detectable [Oii] emission as a
function of their LW. This plot shows that the emission lines of most galaxies are
resolved, but that all LWs are small: σv < 100 km s−1 . There are weak correlations
between the observed LW (σv in Fig. 4) and other global galaxy properties. Firstly,
the LWs appear to be somewhat larger for galaxies of greater [Oii] emission line
luminosity. Secondly, the median LW tends to increase with increasing continuum
12
Rix, Guhathakurta, Col less, Ing
(B band) luminosity. The statistical significance of the trends and a comparison with
corresponding measurements in the local gal.axy population will be given in Sec. 3
and 4.
The [Oii] doublet flux ratio, R[OII] , for most galaxies is found to be near the low
density value of 1.45 (for ne <∼ 102.5 cm−3 ; see Osterbrock 1989, p. 134), consistent
with the line ratios found in Hii regions. Five of the 24 galaxies in our sample have
doublet line flux ratios, R[OII] << 1.32, indicating gas densities significantly (≥ 2σ)
in excess of 100 cm−3 . These ob jects are shown by the open symbols in Figure 4.
The [Oii] flux in these may arise predominantly from an active nucleus, as the LWs
and inferred gas densities are comparable to those of LINERS (Ho, Filippenko and
Sargent, 1993). In this small, but non-negligible fraction of our blue galaxy sample
(5/24), the large emission line EW may not indicate vigorous star formation, but
rather an AGN. This is in agreement with the spectroscopic results of Tresse et al.
(1996) who find that the integrated emission line flux ratios for ∼ 20% of blue field
galaxies are inconsistent with standard Hii region spectra, and are similar to the
ratios found in LINERs. Since we are interested in what the [Oii] linewidths might
tell us about the global gas kinematics on kiloparsec scales, we exclude these five
ob jects from the subsequent analysis.
3 ANALYSIS
In this section, we describe a set of steps that are essential for understanding what our
LW measurements might be telling us about luminosity evolution in galaxies. First,
we show that the line profiles must be broadened by global orbital motion (rotation)
of the ionized gas in the galaxy; this is done by showing that the observed LWs are
much larger than expected from the turbulent motions in individual star forming
complexes. Second, we quantify the bias in the measurement of the rotation speed vc
from the LW measurement by using the observed properties of nearby small galaxies,
the potential local counterparts of the distant galaxies in our sample. For a null-
Linewidths of Distant Field Galaxies
13
hypothesis prediction we use Fabry–Perot (FP) datacubes of these nearby galaxies to
simulate the expected distribution of measured LWs, σv at given vc for these galaxies
if they were observed at z ∼ 0.25 with our AAT/AUTOFIB fibre setup. Third, we
devise a maximum likelihood test to determine what range of linewidth–luminosity–
colour relations are consistent with the data.
3.1 Do the Measured Linewidths Reflect Turbulent Gas Motions?
3.1.1 Relation between Linewidth and Emission Line Luminosity
Melnick et al. (1989) have shown that there is a well determined relation between the
Hβ line luminosity, LHβ , and the mean (Gaussian) LW, σL , for giant Hii regions and
“Hii galaxies”. Presumably, these LWs arise from turbulent motions within the Hii
regions, powered by star formation activity and the resulting supernovae. Melnick et
al. find that this relation takes the form:
log(σL ) = log(σ40.5 ) + η log(LHβ /L40.5 )
(3)
where σ40.5 is the mean LW at LHβ = L40.5 ≡ 1040.5 ergs s−1 cm−2 and the slope of
the relation is η = 0.2. This relation holds over the range 38.5 ≤ log(LHβ ) ≤ 42, with
a remarkably low scatter of only ∆L ∼ 0.06 in log(σL ). The emission line broadening
due to turbulent motion is expected to be very similar for different emission lines.
To check whether our L[OII]–σv data for distant galaxies are consistent with the local
LHβ –σv relationship, we use the observed mean line flux ratio for field galaxies at
hzi ∼ 0.25 (Tresse et al. 1996): hL[OII]/LHβ i ∼ 4.5, with a scatter of a factor of two.
Applying this mean line flux ratio and adjusting the Melnick et al. (1989) fit to H0 =
70 km s−1 Mpc−1 , we expect log(σ40.5 ) = 1.28 for the [Oii] line—i.e., σL ∼ 19 km s−1
for a galaxy with an [Oii] emission line luminosity of L[OII] = 1040.5 ergs s−1 cm−2 .
14
Rix, Guhathakurta, Col less, Ing
3.1.2 Likelihood Analysis
We apply a maximum likelihood test to our galaxy sample in order to determine the
best fit parameters [log(σ40.5 ), η , ∆L ] and their associated errors. The relation given
ppred
L (σv ) =
in Eq. (3), with an intrinsic scatter ∆L in log(σL ), predicts a LW distribution:
[log(σv ) − log(σL )]2
! ,
exp −
1
√2π∆L
2(∆L )2
where σL is the mean velocity dispersion predicted for an “Hii galaxy” with [Oii] line
(4)
pobs (σv ) =
, is approximated by
p[log(L[OII] ), σobs
v
luminosity, L[OII] . On the other hand, the probability distribution of the (true) LW,
and its error, ∆obs
σv , given a measured value of σobs
v
v
v )2
(σv − σobs
v )2 ! .
exp −
1
√2π∆obs
2(∆obs
v
For any parameter set [log(σ40.5), η , ∆L ], the probability of making the observation
[log(L[OII] ), σobs
v
], is given by
] = Z ∞
0
The likelihood of a L[OII]–σv relation, given N observations, is then
N
)i .
loghp(log(L[OII] ), σobs
Lhlog(σ40.5 ), η , ∆L i =
Xi=1
v
The best parameters are determined by maximizing the likelihood in Eq. (13) and
their confidence intervals are derived from the distribution of 2(Lmax − L) (cf. Wilks
1962), which is asymptotically distributed as a χ2 distribution of n degrees of freedom
pobs (σ) ppred
L (σ) dσ .
(5)
(6)
(7)
and where n is the number of parameters considered simultaneously. The maximum
likelihood analysis yields the following results: (a) a well determined zero point of
the relation, log(σ40.5 ) = 1.62 ± 0.05 or σ40.5 = 42 ± 3 km s−1 , which is larger by
7σ than the value in the Melnick et al. (1989) relation [log(σ40.5 ) = 1.28 or σ40.5 =
19 km s−1 ];
(b) a slope η = 0.19 ± 0.08, (c) a scatter ∆L = 0.1+0.06
−0.10 , indicating that
most of the observed scatter is attributable to measurement error. Figure 5 shows
the data points, the best fit (bold solid line) and its uncertainties (dot dashed lines),
and the Melnick et al. relation (dotted line).
Linewidths of Distant Field Galaxies
15
A direct likelihood ratio test shows that our best fit relation is inconsistent with
the relation found by Melnick et al. (1989) at the ≥ 99.99% level. At an [Oii] line
luminosity of L[OII] = 3 × 1040 ergs s−1 cm−2 , the mean LW observed in our distant
galaxy sample is a factor of two higher than that predicted by the σv –LHβ relation
for local Hii galaxies. Put differently, local Hii galaxies, whose turbulent LWs are
comparable to those observed for the distant galaxies should have 50 times higher
line luminosities. This suggests that the observed [Oii] LWs in our distant galaxy
sample are not determined by the turbulent motions within star forming regions,
but rather reflect the global rotational motion of a disk of ionized gas, as they do in
samples of local spiral galaxies.
3.2 Relating the Measured Linewidth to the Galaxy Rotation Speed
We calculate what distribution of measured linewidths, p(σv vc), we should expect
for typical local galaxies of a given rotation velocity, vc , if we observed them at
z = 0.25 with AAT/AUTOFIB. Such a mapping must account for the lack of spatial
information in fibre spectroscopy and the effects of:
• A finite fibre aperture (D = 2 .′′1) and fibre pointing errors of ∼ 0 .′′3 (a conservative
assumption as engineering tests indicate a 0 .′′2 pointing error).
• The shape of the rotation curve [vc (R) 6= constant]
• Clumpy and asymmetric spatial distribution of the ionized gas emissivity
• The absence of information about the disk inclination, i
• Seeing (FWHM ∼ 1 .′′7)
• Fitting a Gaussian model line profile even though the “true” [Oii] line profiles of
galaxies may have a variety of different shapes
3.2.1 Construction of Galaxy Kinematical Models
As a first step, we construct a complete, two-dimensional model for the gas kinemat-
ics of each of three well-resolved local galaxies, ESO 215-G39, ESO 437-G34, and
ESO 323-G73. The model specifies the mean velocity at every radius R and the line
16
Rix, Guhathakurta, Col less, Ing
emissivity at every point (R, θ) in the disk plane. “Reduced” FP data for the three
local galaxies were kindly provided by T. Williams and P. Palunas. The basic prop-
erties of these galaxies are listed in Table 2. The galaxies were chosen because they
cover a range in absolute magnitude and rest-frame colour similar to the members
of the distant sample. The reduced FP data for each galaxy consist of: (a) the Hα
“flux image”, I (x, y ), as a function of position in the plane of the sky, and (b) the
velocity field, vpro j (x, y ), at points of sufficient line flux [I (x, y ) ≈ 0 at many points
in the image]. The velocity field data are too sparse to allow the usual tilted ring
fit; we instead resort to a more restricted model, one in which the galaxy is assumed
to be axisymmetric and coplanar, and to have a rotation curve which can be well
approximated by the parametric form:
vrot (R) ≡ vc (1 + x)β (1 + x−γ )−1/γ
where x ≡ R/R0 . Despite the awkward appearance of the above functional form, each
parameter has a simple interpretation: vc is the velocity scale, R0 is the “turnover”
(8)
,
radius (or core radius), γ determines the sharpness of the turnover (the higher the
value of γ , the sharper the turnover), and the index β specifies the power-law behavior
of the curve at large radii. This form of the rotation curve has a linear rise at small
radii with a slope: ∂vrot /∂R ≈ vc/R0 . In practice, the rotation curves are nearly flat
at large radii (β < 0.1), and we set β = 0 in our model fits in order to have a well
defined velocity scale.
In addition to the four rotation curve parameters, we fit the galaxy’s center
(x0 , y0), mean recession velocity, inclination, and ma jor axis position angle. All
nine parameters describing the velocity field are well constrained by the data for
each of the three local galaxies, and these can be used to depro ject the Hα image,
I (x, y ) → I (R, θ). Some of these parameters are listed in Table 2. The best fit rota-
tion curve and azimuthally averaged Hα flux distribution are shown in Figure 6. This
figure also shows the size of a 2 .′′1 fibre aperture if these local galaxies were observed
Linewidths of Distant Field Galaxies
17
at z = 0.25 (D = 7.6 kpc). The aperture extends to twice the turnover radius R0 for
two of the galaxies, while Rfibre ≈ R0 for the third.
3.2.2 Projection of Kinematical Models
The depro jected models of the three local galaxies, consisting of the ionized flux im-
age, I (R, θ) and the rotation curve, vrot (R), are pro jected to simulate the appearance
of a z = 0.25 galaxy viewed from any arbitrary direction. To calculate the probability
distribution, p(σv vc), of measuring a width σv in a galaxy with a characteristic rota-
tion speed vc , we need to specify the distribution of disk inclination angles, p[cos (i)].
A magnitude limited sample of dust free disk galaxies is expected to have a uniform
distribution in cos(i).
For nearby spiral galaxies, however, it is well established that an inclination-
dependent “internal extinction correction” must be made in calculating the true
luminosity from the observed brightness, to account for dust within the disk of the
galaxy. The magnitude of this correction is still under some debate (cf. Rix 1995),
but the B band correction given in the Third Reference Catalog (de Vaucouleurs et
al. 1991, hereafter RC3) is likely to be an upper limit: AB (i) = 1.5 min[log10 (a/b), 1],
where the axis ratio b/a equals cos(i) for an infinitely thin circular disk. Galaxies at a
given inclination i with “observed” (extincted) luminosities in the interval [L, L + dL]
have “true” (unextincted) luminosities in the interval [eAB (i)L, eAB (i) (L + dL)]. Since
the value of AB (i) increases for more edge-on disks, a magnitude limited sample may
be skewed towards face-on galaxies. For sub-L∗
B galaxies, however, this effect is small
if the faint-end power law exponent of the luminosity function, α, is close to −1. If
the luminosity function is a power-law with α = −1 there are equally many galaxies
in the interval [L, L + dL] and [γL, γ (L + dL)]. Specifically, if we adopt the faint-end
slope for the local galaxy luminosity function α ≈ −1.2 (Efstathiou et al. 1988), the
inclination distribution is only slightly skewed, with hcos(i) = 0.54i, compared to
hcos(i) = 0.5i for the dust-free case. Further, the average velocity reduction due to
pro jection, hsin(i)i differs by only 6% from the dust-free case.
18
Rix, Guhathakurta, Col less, Ing
The presence of dust also implies that a given apparent magnitude corresponds to
a higher intrinsic luminosity than if the galaxy were free of dust. The average value
of the internal extinction is hAB i ≈ 0.4 for the skewed cos(i) distribution (compared
to hAB i = 0.6 for a uniform distribution), and we adopt this value in the subsequent
comparison of our distant galaxy sample to local samples (Sec. 4).
We simulate the AAT/AUTOFIB fibre observations by drawing an inclination
from p[cos (i)] and a random ma jor axis position angle ϕma j and by calculating the
pro jected velocity, vpro j , at each point in the sky plane (x, y ) of the pro jected model of
a local galaxy. To account for turbulent dispersion intrinsic to individual Hii regions,
the line profile at each point is assumed to be a Gaussian with a velocity dispersion
of 15 km s−1 .
Each pixel is given weight in proportion to its Hα emission line flux and the
seeing is simulated by convolving each flux point (spatially) with a circular Gaussian
of FWHM = 1 .′′7. The emissivity weighted line profile is then obtained by integrating
over the area of the (pro jected) galaxy model corresponding to the 2 .′′1 (diameter)
aperture of the AUTOFIB fibres. This aperture corresponds to a radius of 3.8 kpc
at a redshift of 0.25. The effect of pointing error is simulated by shifting the center
of the aperture (over which the integration is carried out) in a random direction,
with the size of the shift drawn from a Gaussian distribution with rms dispersion
0 .′′3. The integrated line profile for a given pair of viewing angles (i, ϕma j) is fitted
to a Gaussian in a least-squares sense; this is a non-trivial step if the true profile
shapes are not Gaussian. Summing the pro jected galaxy model (of given vc ) over
all orientation angles (i, ϕma j ) produces the probability, pMC (σv vc), of measuring a
dispersion σv with our fibre setup.
The probability distribution pMC (σv vc ) differs for the three galaxies because of
differences in the shape of the rotation curve, in the spatial distribution of the emission
line flux from ionized gas, and, most importantly, in the rotation velocity scale vc .
Also, we need to predict pMC (σv vc) for any vc in a certain range, even though we only
have a small number of calibrating galaxies. We do this by assuming that the shape
Linewidths of Distant Field Galaxies
19
of pMC is constant and that the distribution scales with vc . Mathematically speaking,
we assume that pMC (σv vc) = p(σv /vc). The probability distributions for the three
local galaxies are combined and the resulting mean p(σ/vc ) is shown in Figure 7 along
with the rms spread between the three curves. Note that this wide range of observed
LWs, σv , at a given vc , would arise even for infinite signal-to-noise observations with
our technique.
The shape of p(σ/vc) deserves some comment, as it has a broad distribution with
a mean value of σv /vc ∼ 0.6. These features arise from a combination of factors.
For an edge-on galaxy with a uniform flux distribution in the plane of the disk, a
perfectly centered aperture that is large enough to include the flat part of the rotation
curve yields an integrated emission line profile is roughly rectangular; the Gaussian
that best fits a rectangle extending between −vc and +vc has a width σv ∼ 0.75 vc .
This explains why p(σv /vc) falls off for σv /vc >∼ 0.8. Any asymmetry in the shape
of the integrated line profile, due to either a non-uniform flux distribution or fibre
centering error, reduces further the fitted value of σv . The finite fibre size and central
concentration of the ionized flux distribution tend to weight the rising part of the
rotation curve at small radii at the expense of flat part at larger radii, and this too
reduces the measured σv relative to vc . The effect of seeing tends to increase the
value of σv /vc since it causes flux from the outer parts of the galaxy (at extreme
velocities) to spill into the fibre aperture and flux from the central part of the galaxy
to spill out of it. The tail at the low end of the distribution (σv /vc <∼ 0.4) arises
from face-on galaxies. For a randomly oriented galaxy sample, the mean reduction in
velocity due to pro jection effects (compared to an edge-on sample) is hsin(i)i ∼ 0.85.
All the above factors—the definition of σv , line profile asymmetries (caused by spatial
inhomogeneities in the ionized flux and by fibre centering errors), central weighting
of the galaxy (due to finite fibre size and centrally peaked ionized flux distribution),
smearing due to seeing, and inclination effects—are of comparable importance, and
all of them except seeing will reduce σv /vc .
While we measure [Oii] for the distant galaxies, the p(σv /vc) was derived from the
20
Rix, Guhathakurta, Col less, Ing
Hα distribution in local galaxies. It is interesting to ask how the spatial distributions
of [Oii] and Hα compare in local galaxies, since that determines the applicability of
the derived p(σv /vc) to the distant galaxy sample. In a study of the star forming
regions in 14 nearby spiral galaxies, Zaritsky et al. (1994) find that the [Oii] flux
is less centrally concentrated than the Hα flux. This is caused by a radial gradient
in the intrinsic Hα/[Oii] fluxes of Hii regions (which, in turn, is related to their
radial metallicity gradient) and in the amount of (selective) dust extinction (both
increase with increasing radius). Both these effects cause our estimate of σv /vc to
be smaller than the “true” value, making our best fit value of the rotation speed, vc
(see Sec. 3.3.2) somewhat of an overestimate. This would imply that the discrepancy
between the mean rotation speed of our distant galaxy sample and that of local
galaxies is larger than we indicate in Sec. 4.
3.3 The Linewidth–Luminosity–Colour Relation
3.3.1 Fitting Procedure
We are now in a position to test whether our observations are consistent with
any linewidth–luminosity (LWL) relation, or more generally, with any linewidth–
luminosity–colour (LWLC) relation of the form:
loghvc (MB , B − R)i = log [vc (−19, 1)] − η (MB + 19) + ζ [(B − R) − 1]
where vc (−19, 1) is the mean circular velocity of a galaxy with absolute blue mag-
nitude MB = −19 and colour B − R = 1, corresponding to our sample median.
The dependence of the mean circular velocity on galaxy luminosity and colour are
(9)
,
specified by the slopes, η and ζ , respectively.
If the LWLC relation has an intrinsic (logarithmic) scatter of ∆v in velocity, then
the probability of finding vc , for a given luminosity and colour, is
exp − (cid:16)log[vc ] − log[vc (MB , B − R)](cid:17)2
2∆2
v
p[vc MB , B − R]
1
√2π∆v
dvc
vc
=
! dvc
vc
.(10)
Linewidths of Distant Field Galaxies
21
As long as ∆v ≪ 1, it matters little whether the scatter is specified in the velocity
or in its logarithm. The predicted probability of measuring a LW σv for a galaxy of
absolute magnitude MB and colour B − R is then
dvc
ppred (σv MB , B − R) = Z ∞
pMC (σv vc ) p[vc MB , B − R]
vc
0
where the averaging and bias introduced by our observing technique (see Sec. 3.2)
are incorporated through pMC(σv vc). As in Sec. 3.1.2, this distribution must then
be convolved with the measurement uncertainties in order to yield the probability of
(11)
,
p(σobs
v
, given an assumed LWLC:
each galaxy observation, σobs
v
MB , B − R) = Z ∞
pobs (σv ) ppred (σv MB , B − R) dσ ,
0
where pobs (σv ) is defined in Eq. (5). The likelihood, L, of such a LWLC given all N
observations is then
(12)
L[vc (−19, 1), η , ζ , ∆v ] =
N
Xi=1
log[p(σobs
v
MB , B − R)]
.
(13)
3.3.2 Likelihood Limits
The most likely values for the above parameters are determined by maximizing the
likelihood in Eq. (13); their confidence intervals are derived from the distribution of
2(Lmax − L) (cf. Wilks 1962). Figure 8 summarizes the result by showing two pro jec-
tions of the likelihood contours, that include the best fit value, in the [vc (−19, 1), η ]
and [vc (−19, 1), ζ ] planes, illustrating the dependence of vc on MB and (B − R),
respectively. The analysis yields best fit values of: vc (−19, 1) = 66 ± 8 km s−1 ,
η = 0.07 ± 0.08, and ζ = 0.28 ± 0.25. The pro jections of the best fit onto the
MB − η and MB − ζ plane are compared to the data in Figure 8. The likelihood limits
of the parameters in the MB − η and MB − ζ planes are shown in Figure 9.
The “zero point” of the LWLC relation is well determined. Note that the internal
dust extinction is included in these models and enters in two ways: (1) it leads to
a 0.4 mag correction for the mean B -band internal extinction for inclined spirals
(Sec. 3.2.2), and (2) it results in a slight skewing of the inclination distribution towards
22
Rix, Guhathakurta, Col less, Ing
face-on galaxies, changing the inferred linewidth by ∼6%; these two effects “cancel”
each other partially. There is some covariance between the zero point and the slope,
η ≡ −∂ log(vc )/∂MB , because the median absolute magnitude of the distant galaxy
sample is somewhat brighter than our reference value, MB = −19.
As the top panel of Figure 9 shows, a wide range of slopes, η ≡ −∂ log(vc )/∂MB ,
is consistent with the data, including η = 0, the case of no luminosity dependence of
vc . This is caused in part by the limited absolute magnitude range of our sample of
galaxies (∆MB ∼ 2 mag) and in part by the breadth of the p(σv /vc) distribution. Note
that the typical LW–luminosity slope observed for local samples, η ∼ 0.12 (Bothun
et al. 1985; Fukugita et al. 1991), is also consistent with our data.
The bottom panel of Figure 9 shows the colour dependence of vc through the
likelihood contours in the [vc (−19, 1), ζ ] plane, at fixed η . The dependence of vc
on B − R, suggested by the bottom panel of Figure 4, is not significant: ζ ≡
∂ log(vc )/∂ (B − R) = 0.28 ± 0.25 (68% confidence limits).
In order to test the sensitivity of these results to the exact shape of the proba-
bility distribution, p(σv /vc ), we have repeated the above likelihood analysis with the
p(σv /vc ) distributions derived separately for each of the three local calibrator galax-
ies. The three probability distributions are roughly similar in shape as shown by the
error bars in Figure 7. The resulting 68% likelihood regions for any given parameter
overlap substantially for these three realizations, indicating that the results for the
three are consistent with one another.
The intrinsic scatter in the LWLC relation of distant galaxies [defined in Eq. (10)]
is only weakly constrained: ∆v < 0.32 (90% upper limit). This may be understood as
follows. The probability distribution, p(σv /vc ), for any given vc is very broad even if
∆v = 0. In the presence of finite intrinsic scatter in the LWLC relation, the resulting
distribution of σv is a convolution of the intrinsic vc spread (Gaussian with dispersion
∆v ) with p(σv /vc ). The results are largely independent of the intrinsic scatter ∆v ,
as long as it is at least a factor of two smaller than the spread in observed LWs,
p(σv /vc ), introduced by our LW measuring technique. This implies that the results
Linewidths of Distant Field Galaxies
23
(including the size of the confidence regions) are insensitive to the value of ∆v ; we
fix this parameter at its best fit value, ∆v = 0.15, in the rest of the analysis. Note,
the intrinsic scatter is also consistent with zero.
4 COMPARING THE LINEWIDTH–LUMINOSITY RELATION FOR
DISTANT VERSUS LOCAL GALAXIES
We now turn to a quantitative comparison of the LWLC in our distant sample to the
properties of present day galaxies. This requires LW measurements in a set of local
galaxies whose B luminosities and colours are similar to those of the distant field
galaxies in our sample. Unfortunately, most galaxy samples targeted in Tully–Fisher
studies are designed for optimal distance estimates, but do not provide an unbiased
estimate of the LWLC relation for a statistically well defined population.
4.1 Relating the HI Linewidth W20 to the Optical Linewidth vc
A large number of local galaxy LW measurements are based on Hi single-dish data,
usually defined to be the width of the Hi line profile at 20% of the peak intensity,
W20 (or equivalent measures such as W50 ). To relate W20 to the optically measured
“circular velocity” vc , it is customary to set:
vc = 0.5 f W20/sin(i) .
(14)
There are two main reasons why the correction factor f differs from unity. First,
turbulent motions broaden the Hi profile. Second, rotation curves are not perfectly flat
at large radii, and the Hα-emitting ionized gas and neutral Hi gas (used to measure vc
and W20 , respectively) may sample different radii. The relative importance of these
two effects is still under debate and we resort to published Hα and Hi studies to
obtain an empirical calibration of the vc - W20 relation. One of the largest samples of
optical rotation curves is available from the survey by Mathewson et al. (1992a). In
their analysis Mathewson et al. (1992b) they find that vc (Hα) ∼ 0.94 (0.5 W50/sin(i))
for galaxies with MB ∼ −19; Rubin et al.(1985) and Courteau (1992) find that on
24
Rix, Guhathakurta, Col less, Ing
average hvci = v ( W50) within a few percent, albeit for more luminous galaxies. If
these results are combined with W50 ≈ 0.9 W20 (e.g.Bothun et al.1985) this implies a
correction factor of f = 0.86 ± 0.03. Schommer et al. (1993) applied the turbulence
correction advocated by Tully & Fouqu´e (1985) to a sample of galaxies for which they
had both optical (Hα) and Hi measurements (corresponding to f = 0.83), but found
that a smaller correction (f = 0.9) was needed to bring the two measurements into
agreement. In summary, these studies indicate that a correction of f = 0.86 ± 0.04
should be applied for galaxies with MB ∼ −19 to go form W20/sin(i) to 2vc .
In the subsequent comparison with local data, we adopt a correction factor of
f = 0.86, which corresponds to correcting W20 downward by 14%.
4.2 Existing Tully–Fisher Studies
We consider two B band Tully–Fisher samples of nearby disk galaxies—those of
Bothun et al. (1985) and Fukugita et al. (1991)—for comparison to our distant galaxy
sample. The member galaxies of both of these samples are distant enough to yield
sufficiently good relative distances relative to our hzi ∼ 0.25 sample, once scaled
to the same Hubble constant. Further, both these Tully–Fisher samples contain a
sizeable number of galaxies with luminosities as low as MB ∼ −18, and thus overlap
in luminosity (and colour) with our sample of distant galaxies.
The B band study of Coma cluster galaxies by Fukugita et al. (1991) yields a mean
rotation velocity, vc(−19) = 102 km s−1 , for galaxies with MB = −19. Bothun et al.
(1985) list B −V colours for their galaxies, allowing us to mimic the bJ −rF < 1.2 color
cut for the distant field galaxy sample. The resulting mean colour of this subsample
is B − V = 0.36, comparable, after K-corrections and color transformations (Fukugita
et al. 1995) to the mean (observed) colour of our hzi ∼ 0.25 galaxy sample. The zero
point of the LWL relation for the colour-restricted Bothun et al. galaxy subset is not
significantly different from that of the whole sample, vc (−19, 1) = 102 km s−1 , and
is identical to Fukugita et al.’s 1991 value.
As the the LWs in these local Tully–Fisher samples were derived from Hi measure-
25
ments, W20 , the quoted value for the “optical” rotation speeds, vc (−19) = 102 km s−1 ,
include the 14% downward correction from Section 4.1. The parameters of the best
Linewidths of Distant Field Galaxies
fit LWL relations for these two local Tully–Fisher samples are shown in Figure 9,
with a correction for W20 → vc .
4.3 Local Linewidth–Luminosity–Colour Relation from RC3
Most nearby Tully–Fisher samples, including Fukugita et al. (1991) and Bothun et al.
(1985), are restricted to a narrow range of Hubble types and usually exclude irregular
galaxies. Our hzi ∼ 0.25 sample, however, has no morphological restriction. Since the
mean vc at a given MB may vary with Hubble type (Rubin et al.1985), the local
Tully–Fisher galaxies may not be directly comparable to the distant ones.
As a remedy, we have selected from the RC3 catalog a set of local galaxies with
measured W20 , ignoring their Hubble type, requiring only:
a. Axis ratios b/a < 0.7, so that the uncertainty in the LW sin−1(i) depro jection
factor is small
b. Recession velocities cz > 1000 km s−1 and angular separations of > 10◦ from the
center of the Virgo cluster, so that our distance estimate, D = cz/H0 , is not
severely affected by peculiar velocities
The resulting RC3 sample includes galaxies as faint as MB = −17 and many galaxies
with −18 > MB > −20, the absolute magnitude range spanned by our hzi ∼ 0.25
sample.
We construct a LWL relation for the blue (B − V < 0.5, hB − V i = 0.41)
and red (B − V > 0.5, hB − V i = 0.66) subsample separately. This colour cut
corresponds to the colour threshold, bJ − rF < 1.2 used to select the distant galaxy
sample. Both blue and red RC3 samples show a clear LWL trend, despite a scatter
of almost a factor of 2 in vc at fixed MB . This scatter appears to be dominated by
disk inclination errors. The zero point of the best fit LWL relation for the blue RC3
sample, vc (−19, blue) = 110 km s−1 , is in reasonable agreement with the zero points
26
Rix, Guhathakurta, Col less, Ing
of the morphologically-restricted Fukugita et al. (1991) and Bothun et al. (1985)
samples. As discussed above, these zero point estimates include a 14% downward
correction from 0.5 W20/sin(i) → vc . The best fit slope of the RC3 blue galaxy LWL
relation, η = 0.11, is also consistent with that of the Fukugita et al. and Bothun et
al. samples and with that of the distant galaxy sample.
The mean rotation velocity of MB = −19 galaxies in the red RC3 sample,
vc (−19, red) = 129 km s−1 , is significantly higher than in the blue sample, although
the LWL slopes of the two samples are identical. This trend in the local vc versus
colour relation corresponds to a slope, ζ = +0.11 (after converting the rest-frame
B − V colour for local galaxies to a redshifted bJ − rF colour at z = 0.25).
The colour dependence of vc in local galaxies has an interesting implication. If
the difference between the mean rotation speeds, vc (−19, 1), of distant and local
galaxies is caused by galaxies being brighter in the past, one also expects them to
be somewhat bluer in the past (due to a younger mix of stars). In other words, the
appropriate local counterpart of a z = 0.25 galaxy with MB = −19 and B − R = 1, is
not only one that is fainter but also one that is redder (B − R > 1). Since the mean vc
increases for redder galaxies locally, the offset in vc between distant and local galaxies
(and the amount of luminosity evolution it implies) would be even larger than we
calculated in Sec. 4.2 had we taken colour evolution into account.
In summary, comparison of our linewidth data on blue field galaxies at hzi ∼ 0.25
with nearby galaxy data leads us to conclude that, at the > 99% confidence level,
the distant galaxy sample does not have rotation speeds as high as that of local
samples with similar photometric properties (B luminosity, B − R colour). Taken at
face value, the best fit rotation speed of blue MB = −19 galaxies, vc (−19, 1), is 35%
smaller (at least 25% smaller at 90% confidence) than the local value. It is likely
that this difference can be attributed to an epoch of increased star-formation, which
lead to a larger luminosity at a given LW. Assuming the local B-band slope for the
LWL relation (η ≈ 0.12), the off-set in vc (−19, 1) correponds to a magnitude offset
of ∆MB = 1.5 mag, with a lower limit of 0.8 mag (99%).
Linewidths of Distant Field Galaxies
27
5 CONCLUSIONS AND FUTURE WORK
We have presented results from an exploratory pro ject to determine the relation
between the kinematic and photometric properties of galaxies at cosmologically sig-
nificant distances and to compare it to the relation found for local samples. The target
of our study is a statistical sample of blue, sub-L∗ field galaxies in the redshift range
0.16 < z < 0.37, for which we have obtained high dispersion spectra to determine
their [Oii] emission LWs, integrated over an aperture of ∼ 7.6 kpc diameter. The [Oii]
LWs are resolved (σ >∼ 30 km s−1), yet the measured dispersions σv are < 100 km s−1
for all sample galaxies. For several reasons we believe that we have detected the
[Oii] emission and have measured LWs for most galaxies within the magnitude, color
and redshift range; the sample is nearly complete. With the imposed colour cuts
the sample constitutes the blue half of the galaxy population in this luminosity and
redshift range.
The primary goal of our pro ject is to test the null hypothesis that these distant
galaxies have the same characteristic circular velocities, vc , as local galaxies of the
same absolute B magnitude and colour. We have used Hα Fabry–Perot data on local
galaxies to mimic our fibre observations, and to thereby calibrate the biases/errors
associated with the measurement of vc from emission LWs. The following main con-
clusions arise from this analysis:
1. At a given emission line luminosity, the observed [Oii] linewidths of distant galax-
ies are too large to be compatible with the relation between line luminosity and
(turbulent) LW observed for nearby giant Hii regions and Hii galaxies. Further,
80% of the our sample members have [Oii] doublet flux ratios that indicate gas
densities of <∼ 100 cm−3 , consistent with Hii region spectra. Hence, the LWs of
most ob jects reflect the orbital motion of ionized gas in the galaxy’s potential
28
Rix, Guhathakurta, Col less, Ing
well. In the remaining 20% of ob jects, the emission line flux may arise from an
active nucleus.
2. We have tested whether distant blue galaxies within a narrow range of absolute
blue luminosity have the same mean vc as local galaxies of the same absolute
magnitude and colour. In order to make a realistic comparison, we have simulated
what emission LWs would be measured at z ∼ 0.25 for local galaxies of a given vc ,
assuming that the shape of the rotation curve and spatial distribution of ionized
gas are comparable in distant and local galaxies. We find that, at a given lumi-
nosity, MB ∼ −19, the distant galaxies have rotation speeds that are about 35%
smaller than expected from photometrically identical, local samples. The LWs of
the two samples are inconsistent at the > 99.9% confidence level.
3. The most likely explanation for the LW offset is that galaxies have undergone
considerable luminosity evolution. Assuming the local B-band slope for the LWL
relation (η ≈ 0.12), the off-set in vc (−19, 1) correponds to a magnitude offset
of ∆MB = 1.5 mag, with a lower limit of 0.8 mag (99%). If spectral evolution
accompanies this luminosity evolution, the magnitude offset may be even larger.
However, at this point it cannot yet be ruled out that the LW differences in the
samples arise, at least in part, from a very different vc →[Oii] mapping. Such
differences could arise from either the shape of their rotation curve, and/or the
spatial distribution of the line emitting gas.
This systematic investigation of low-luminosity, blue field galaxies, complements
corresponding studies of massive, red cluster galaxies. Franx (1993) and van Dokkum
and Franx (1996) studied the fundamental plane of early type galaxies in two clusters
at z = 0.19 and z = 0.4. They found that galaxies of a given mass scale were brighter
in the past by an amount that is consistent with a formation at high z and subsequent
“passive evolution”. Also, Lilly et al.(1996) find that the luminosity function evolves
most strongly for low-luminosity blue galaxies. The present data indicate that this
evolution of the luminosity function can be traced to the mass-to-light ratio evolution
Linewidths of Distant Field Galaxies
29
of individual galaxies. However, we infer values of vc for galaxies of MB = −19 which
are not as low as predicted for some dwarf starburst scenarios (e.g.Babul and Rees,
1992; Babul and Ferguson, 1996).
The present study represents only one step towards understanding the internal kine-
matics of distant field galaxies. Follow-up should address the following questions:
a. Is there evern more direct evidence that emission LWs reflect ordered rotation of
ionized gas within these galaxies? We are currently pursuing this question with the
help of Fabry–Perot imaging datacubes for a small sample of distant field galaxies
(Ing et al. 1996; see also Vogt et al. 1993, 1996).
b. What is the spatial distribution of the emission line flux in these distant galaxies?
Is it the same as for local galaxies, or are the observed small LWs in part due to
a much more centrally concentrated flux distribution? A direct answer to these
questions would require high angular resolution, narrow band images tuned to the
redshifted [Oii] line.
c. Can the scatter in the linewidth–luminosity relation be reduced if we correct for
galaxy inclinations? A reduction in scatter is expected if, as we argue, the LW
is dominated by gas motions in a disk. Inclinations of sufficient accuracy can be
obtained from HST images.
d. What is the slope of the linewidth–luminosity relation for distant galaxies when
we include a wider range of absolute magnitudes? We are analyzing high disper-
sion spectroscopic data obtained with the AAT/AUTOFIB instrument of nearly
100 galaxies spanning a broader range of MB and colour than the sample presented
here.
e. Is there a similar offset in the LWL relation when the luminosity is measured in a
different bandpass? The amount of luminosity evolution is expected to be smaller
in the rest-frame I band than in B . Near infrared (H -band) photometry is being
carried out to test this hypothesis.
30
Rix, Guhathakurta, Col less, Ing
We are grateful to Steve Lee and Raylee Stathakis for assistance with the AUT-
OFIB observations. We would like to thank Sandy Faber and Stephane Courteau for
helpful suggestions.
Linewidths of Distant Field Galaxies
31
REFERENCES
Babul, A., & Rees, M. J. 1992, MNRAS, 255, 346
Babul, A., & Ferguson, H., 1996, ApJ, 458, 100
Bothun, G. D., Aaronson, M., Schommer, B., Mould, J., Huchra, J. and Sullivan, W., 1985, ApJS, 57, 423
Broadhurst, T. J., Ellis, R. S., & Shanks, T. 1988, MNRAS, 235, 827
Broadhurst, T. J., Ellis, R. S., & Glazebrook, K. 1992, Nature, 355, 55
Cole, S., Aragon-Salamanca, A., Frenk, C., Navarro, J., & Zepf, S. 1994, MNRAS, 271, 781
Colless, M. M., Ellis, R. S., Taylor, K., & Hook, R. 1990, MNRAS, 244, 408 (CETH)
Colless, M. M., Ellis, R. S., Broadhurst, T. J., Taylor, K., Peterson, B. A. 1993a, MNRAS, 261, 19
Colless, M. M., Schade, D., Broadhurst, T. J., & Ellis, R. S. 1993b, MNRAS, 267, 1108
Courteau, S., 1992, Ph.D. Thesis, UC Santa Cruz
Cowie, L. L., Gardner, J.P., Hu, E. M., Songaila, A., Hodapp, K. W. and Wainscoat, R. J., 1994, ApJ, 434, 114.
Efstathiou, G., Ellis, R. S., & Peterson, B. A. 1988, MNRAS, 232, 431
Franx, M., 1993, PASP, 105, 1058
van Dokkum, P. and Franx, M., 1996, MNRAS, in press
Fukugita, M., Okamura, S., Tarusawa, S., Williams, B. and Rood, H. 1991, ApJ, 376, 8
Fukugita, M., Shimasaku, K. and Ichikawa, T., 1995, PASP, 107, 945.
Gardner, J.P., Cowie, L. L., & Wainscoat, R. J. 1993, ApJ, 415, L9
Gronwall, C., & Koo, D. C. 1995, ApJ, 440, L1
Guzman, R. et al., 1995, ApJ, 460, L5
Ho, L., Filippenko, A. and Sargent, W., 1993, ApJ, 417, 63
Ing, K., Guhathakurta, P., Rix, H.-W., Williams, T. B., & Colless, M. M. 1996, ApJ, in preparation
Guiderdoni, B., & Rocca-Volmerange, B. 1990, A&A, 227, 362
Jones et al. 1991, MNRAS, 249, 481
Kaufmann, G., Guiderdoni, B., & White, S. D. M. 1994, MNRAS, 267, 981
Koo, D. C., & Kron, R. G. 1992, ARAA, 30, 613
Koo, D. C., et al., 1995, ApJ, 440, L49
Kron, R. G. 1980, ApJS, 43, 305
Lilly, S. J., Cowie, L. L., & Gardner, J. P. 1991, ApJ, 369, 79
Lilly, S. J. et al.1995, ApJ, 455, 108
Mathewson, D., Ford, V., & Buchhorn, M. 1992a, ApJS, 81
Mathewson, D., Ford, V., & Buchhorn, M. 1992b, ApJL, 5
Melnick, J., Terlevich, R., & Moles, M. 1989, MNRAS, 235, 297
Osterbrock, D., 1989 Astrophysics of Gaseous Nebulae, University Science Books, Mill Valley
Peebles, P. J. E. 1993, Principles of Physical Cosmology (Princeton University Press, Princeton)
Rix, H.-W. & White, S. D. M. 1992, MNRAS, 254, 389
Rix, H.-W. 1995, in The Opacity of Spiral Disks, Ed. J. Davies and D. Burstein, Kluwer, Dortrecht.
Rix, H.-W., M. Colless and P. Guhathakurta, 1995, in New Light on Galaxy Evolution, IAU Symp. 171, Ed. R.
Bender and R. Davies, Kluwer
Rubin, V., Burstein, D., Ford, W. and Thonnard, N., 1985, ApJ, 289, 89
32
Rix, Guhathakurta, Col less, Ing
Schommer, R. A., Bothun, G. D., Williams, T. B., & Mould, J. R. 1993, AJ, 105, 97
Thonnard, N. 1983, in Besancson Symp. on Internal Kinematics of Galaxies (Reidel)
Tinsley, B. M. 1972, ApJ, 178, 319
Tresse, L. et al., 1996, astro-ph/9604028
Tully, R. B., & Fouqu´e, P. 1985, ApJS, 58, 67
Tyson, J. A. 1988, AJ, 96, 1
de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Jr., Buta, R. J., Paturel, G., & Fouqu´e, P. 1991, Third
Reference Catalogue of Bright Galaxies (Springer, New York) (RC3)
Vogt, N. P., Herter, T., Haynes, M and Courteau, S. 1993, ApJ, 415, L95
Wardle, J. F. C., & Kronenberg, P. P. 1974, Ap.J., 194, 249
Wilks, S. S. 1962, Mathematical Statistics (Wiley & Sons, New York)
Zaritsky, D., Kennicutt, R. C., Jr., & Huchra, J. P. 1994, ApJ, 420, 87
Table 1. Summary of the Observational Results
R.A. (1950.0) Dec.
mb
b − r
1.07
Linewidths of Distant Field Galaxies
33
z
σa
Mb
b LOII
EW
ROII
21.83
0.28236
0.18024
0.15934
0.21472
0.20723
0.34462
0.23491
0.34455
0.21367
0.26532
0.16399
21.36
21.40
21.25
21.69
0 55 41.91 -27 53 38.2
21.41
0 55 15.69 -27 49 56.6
21.69
0 55 54.30 -27 41 59.0
21.60
0 55 34.02 -27 35 09.5
21.32
0 55 44.01 -27 40 42.4
21.52
0.35
0.92
0.71
0.34
0.52
0.77
0.83
0.80
1.08
0 55 28.46 -27 35 50.5
1.16
1.07
0.96
0.92
0 54 12.55 -28 07 32.3
0 54 27.40 -28 02 31.9
0 54 46.57 -28 00 56.6
0 55 27.81 -28 00 48.6
0 55 20.22 -28 00 29.2
0 55 30.23 -27 57 48.4
60 ± 08
58 ± 05
33 ± 23
62 ± 11
41 ± 16
11 ± 14
46 ± 08
58 ± 19
18 ± 18
34 ± 12
28 ± 21
53 ± 09
75 ± 24
78 ± 15
32 ± 19
27 ± 18
36 ± 11
67 ± 27
45 ± 58
52 ± 05
62 ± 18
86 ± 17
51 ± 15
31 ± 10
The error listed is the geometric mean of the upper and lower error bar.
0 53 53.12 -28 04 53.7
0 53 48.90 -27 55 59.3
0 54 09.32 -27 54 49.2
0 54 44.74 -27 52 34.8
0 54 04.35 -27 48 44.0
0 54 06.74 -27 51 14.0
0 53 48.42 -27 44 59.4
0 54 19.62 -27 48 08.4
0 54 05.09 -27 45 08.8
0.81
1.01
1.05
1.16
1.09
1.18
0.94
1.04
0 54 30.99 -27 49 18.0
21.73
0 53 59.66 -27 37 45.1
21.42
0 54 33.00 -27 45 05.7
21.68
21.94
21.38
21.87
21.65
21.35
21.39
21.79
21.47
0.32507
0.23578
0.21372
0.24532
0.28742
0.33404
0.23972
0.21635
0.21405
0.27436
0.21332
1.17
0.92
0.23260
0.29622
21.75
21.39
21.32
-19.29
-20.36
-18.84
-18.77
-18.58
-20.27
-19.25
-18.51
-17.84
-19.69
-17.97
-18.96
-19.57
-19.24
-18.57
-19.57
-18.49
-19.08
-19.35
-19.04
-19.78
-19.86
-19.16
-19.57
66
65
12
29
22
15
38
14
8
20
19
26
38
33
24
20
39
15
12
61
19
30
24
34
41.1
41.5
40.3
40.6
40.4
40.8
40.9
40.2
39.7
40.7
40.1
40.6
40.9
40.8
40.4
40.7
40.6
40.4
40.4
41.0
40.7
40.9
40.6
40.9
1.36 ± .10
1.26 ± .06
0.62 ± .15∗
1.32 ± .12
1.55 ± .17
0.94 ± .15∗
1.48 ± .10
1.53 ± .22
1.26 ± .34
1.57 ± .16
1.10 ± .16
1.40 ± .11
1.51 ± .21
1.34 ± .12
1.66 ± .22
1.45 ± .16
1.55 ± .15
0.50 ± .38∗
0.79 ± .21∗
1.26 ± .07
1.34 ± .20
0.58 ± .14∗
1.08 ± .13
1.05 ± .12
a
b
The EW has been estimated from the observed [Oii]line flux, the instrument efficiency and the bj magnitude of
the galaxy.
c
The five galaxies where the line ratio is significantly (2σ) below 1.32 have been labelled with an asterisk. Their line
ratios
indicate gas densities in excess of 100cm−3 , and may arise from an AGN, rather than HI regions in the galaxies’
disks.
34
Rix, Guhathakurta, Col less, Ing
Table 2. Local Galaxies for Establishing p(σvc )
Name
ESO 215-G39
v [km/s] MB B − R vcirc R0 [kpc]
1.88
153
0.8
-19.8
4335
ESO 323-G73
5016
-19.9
ESO 437-G43
3801
-17.7
1.0
0.8
157
104
1.83
3.50
i [◦ ]
γ
47
48
53
2.2
2.3
3.7
|
astro-ph/0607015 | 1 | 0607 | 2006-07-03T04:29:30 | Chemical Abundances in the Secondary Star of the Black Hole Binary V4641 Sagittarii (SAX J1819.3-2525) | [
"astro-ph"
] | We report on detailed spectroscopic studies performed for the secondary star in the black hole binary (micro-quasar) V4641 Sgr in order to examine its surface chemical composition and to see if its surface shows any signature of pollution by ejecta from a supernova explosion. High-resolution spectra of V4641 Sgr observed in the quiescent state in the blue-visual region are compared with those of the two bright well-studied B9 stars (14 Cyg and $\nu$ Cap) observed with the same instrument. The effective temperature of V4641 Sgr (10500 $\pm$ 200 K) is estimated from the strengths of He~{\sc i} lines, while its rotational velocity, $\it v$ sin $\it i$ (95 $\pm$ 10 km s${}^{-1}$), is estimated from the profile of the Mg~{\sc ii} line at 4481 \AA. We obtain abundances of 10 elements and find definite over-abundances of N (by 0.8 dex or more) and Na (by 0.8 dex) in V4641 Sgr. From line-by-line comparisons of eight other elements (C, O, Mg, Al, Si, Ti, Cr, and Fe) between V4641 Sgr and the two reference stars, we conclude that there is no apparent difference in the abundances of these elements between V4641 Sgr and the two normal late B-type stars, which have been reported to have solar abundances. An evolutionary model of a massive close binary system has been constructed to explain the abundances observed in V4641 Sgr. The model suggests that the progenitor of the black hole forming supernova was as massive as ~ 35 Msun on the main-sequence and, after becoming a ~ 10 Msun He star, underwent "dark" explosion which ejected only N and Na-rich outer layer of the He star without radioactive $^{56}$Ni. | astro-ph | astro-ph | PASJ: Publ. Astron. Soc. Japan , 1 -- ??,
c(cid:13) 2018. Astronomical Society of Japan.
6
0
0
2
l
u
J
3
1
v
5
1
0
7
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Chemical Abundances in the Secondary Star of the Black Hole Binary
V4641 Sagittarii (SAX J1819.3 -- 2525) ∗
Kozo Sadakane,1 Akira Arai,1 Wako Aoki,2 Nobuo Arimoto,2 Masahide Takada-Hidai,3 Takashi Ohnishi,4
Akito Tajitsu,5 Timothy C. Beers,6 Nobuyuki Iwamoto,7 Nozomu Tominaga,8 Hideyuki Umeda,8
Keiichi Maeda,9 and Ken'ichi Nomoto8
1Astronomical Institute, Osaka Kyoiku University, Asahigaoka, Kashiwara, Osaka 582-8582
2National Astronomical Observatory, 2-21-1 Osawa, Mitaka, Tokyo 181-8588
3Liberal Arts Education Center, Tokai University, 1117 Kitakaname, Hiratsuka, Kanagawa 259-1292
4Nagoya Science Museum, 2-17-1, Sakae, Naka-ku, Nagoya, Aichi 460-0008
5Subaru Telescope, National Astronomical Observatory of Japan, 650 North A'ohoku Place,
6Department of Physics and Astronomy and Joint Institute for Nuclear Astrophysics (JINA),
Michigan State University, East Lansing, MI 48824, USA
7Nuclear Data Center, Japan Atomic Energy Agency, Ibaraki 319-1195
Hilo, HI 96720, USA
8Department of Astronomy, School of Science, University of Tokyo, Bunkyo-ku, Tokyo 113-0033
9Department of Earth Science and Astronomy, College of Arts and Sciences, University of Tokyo, Tokyo 153-8902
[email protected]
(Received 2006 January 12; accepted 2006 April 12)
Abstract
We report on detailed spectroscopic studies performed for the secondary star in the black hole binary
(micro-quasar) V4641 Sgr in order to examine its surface chemical composition and to see if its surface
shows any signature of pollution by ejecta from a supernova explosion. High-resolution spectra of V4641
Sgr observed in the quiescent state in the blue-visual region are compared with those of the two bright
well-studied B9 stars (14 Cyg and ν Cap) observed with the same instrument. The effective temperature
of V4641 Sgr (10500 ± 200 K) is estimated from the strengths of He i lines, while its rotational velocity,
v sin i (95 ± 10 km s−1), is estimated from the profile of the Mg ii line at 4481 A. We obtain abundances
of 10 elements and find definite over -- abundances of N (by 0.8 dex or more) and Na (by 0.8 dex) in V4641
Sgr. From line-by-line comparisons of eight other elements (C, O, Mg, Al, Si, Ti, Cr, and Fe) between
V4641 Sgr and the two reference stars, we conclude that there is no apparent difference in the abundances
of these elements between V4641 Sgr and the two normal late B-type stars, which have been reported to
have solar abundances. An evolutionary model of a massive close binary system has been constructed to
explain the abundances observed in V4641 Sgr. The model suggests that the progenitor of the black hole
forming supernova was as massive as ∼ 35M⊙ on the main-sequence and, after becoming a ∼ 10M⊙ He
star, underwent "dark" explosion which ejected only N and Na-rich outer layer of the He star without
radioactive 56Ni.
Key words: stars:abundances -- stars:binaries -- stars:
individual (V4641 Sagittarii) -- stars: micro-
quasars -- stars: X-rays
1.
Introduction
The object V4641 Sgr is an X-ray binary with an orbital
period of 2.817 d, containing a primary (a black hole) of
∼ 9.6 M⊙ and a secondary star of 5 -- 8 M⊙(Orosz et al.
2001). Extensive photometric observations were carried
out by Goranskij et al. (2003), who reported a photomet-
ric period of 2.81728 d, and V and R bands light curves
which were dominated by ellipsoidal variations from the
secondary. They demonstrated that the surface temper-
ature of the secondary is non -- uniform from the observed
∗ Based on data collected at the Subaru Telescope, which is oper-
ated by the National Astronomical Observatory of Japan.
color variation. They observed a depression in the red
wing of the Hα line just before the black hole's inferior
conjunction, and interpreted this as being due to absorp-
tion by the rarefied gas disk around the black hole.
The source exhibited two rapid and bright X-ray flares
around 1999 September 15 (Smith et al. 1999). Just prior
to the giant X-ray flare, the source showed an increase in
the variability (Kato et al. 1999). About six days before
the X-ray flare, the source exhibited a ∼ 1 mag modula-
tion with a period of ∼ 2.5 d. After reaching the peak
brightness (8.8 mag in V ), it decayed rapidly to its mean
quiescent level within two days. Since then, the source
exhibited several outbursts with very rapid fluctuations
in the optical band (Uemura et al. 2002a, 2002b, 2004a,
2
2004b, 2005).
Orosz et al. (2001) carried out extensive spectroscopic
observations of V4641 Sgr in 1999 and obtained orbital
parameters of the system. The spectral type of the sec-
ondary star was assigned to be B9 III. They determined
the rotational velocity, v sin i, to be 123 ± 4 km s−1,
and performed spectroscopic analyses of the secondary
star using moderate-resolution spectra to estimate the
abundances of several light elements. They found over --
abundances of several elements including N, O, Mg, Ca,
and Ti, with a solar abundance of Si. Although they
note that better data (e.g., high-resolution spectra) or bet-
ter models are needed before establishing the abundance
anomalies, their results on (possible) over -- abundances of
light elements are very interesting because chemical abun-
dances of the secondary star are expected to provide di-
rect information on the products of nucleosynthesis from
supernova explosions of massive stars.
Information on
abundances is also expected to constrain many phisical
parameters that are involved in supernova explosion mod-
els. They include the mass cut, the amount of fall-back
matter, possible mixing, and explosion energies and ge-
ometries.
Israelian et al. (1999) reported over -- abundances of O,
Mg, Si, S, and Ti, but a solar abundance of Fe, in the
eclipsing low -- mass X-ray binary GRO J1655-40 (Nova
Scorpii 1994). Based on this observation and using a va-
riety of supernova models, Podsiadlowski et al.
(2002)
showed that the best fits can be obtained for He star
masses of 10 -- 16 M⊙, where spherical hypernova mod-
els are generally favored over standard supernova ones.
Over -- abundances of Al, Ti, and Ni are reported in the
black hole binary A0620 -- 00 by Gonz´alez Hern´andez et
al. (2004). They discussed a possible scenario of pollu-
tion from a supernova. Gonz´alez Hern´andez et al. (2005)
found enhanced abundances of Ti, Fe, and Ni in the neu-
tron star binary Cen X-4, and showed that these apparent
anomalies can be explained if the secondary star captured
a significant amount of matter ejected from a spherically
symmetric supernova explosion of a 4 M⊙ He core pro-
genitor.
In order to clarify the possible abundance anomalies re-
ported in V4641 Sgr, we carried out high-resolution spec-
troscopic observations using the Subaru telescope and per-
formed comparative abundance analyses using two well-
studied reference stars that are reported to have solar
abundances.
2. Observational Data
Spectroscopic observations of V4641 Sgr were carried
out with the Subaru Telescope using the High Dispersion
Spectrograph (HDS) on 2005 May 21 and June 19 (UT).
Our observations were made during its quiescent state,
though the object has been reported to exhibit an out-
burst on 2005 June 24.371 (UT) (M. Uemura, private com-
munication). Data for two reference stars [14 Cyg = HD
185872 (B9 III) and ν Cap = HD193432 (B9.5 V)] were
obtained on 2005 May 20 and June 19, respectively, using
K. Sadakane et al.
[Vol. ,
the same instrumental setup. Technical details and the
performance of the spectrograph are described in Noguchi
et al. (2002). We used a slit width of 1.′′0 (0.5 mm) and a
2x2 binning mode, which enabled us to achieve a nominal
spectral resolving power of about R = 45000 with a 3.5
pixel sampling. Our observations covered the wavelength
region from 4050 A to 6760 A with a gap between 5340 --
5450 A. The exposure times for V4641 Sgr were 4400 sec
and 5400 sec on May 21 and June 19, respectively. For
flat-fielding of the CCD data, we obtained Halogen lamp
exposures (flat images) with the same setup as that for
the object frames.
The reduction of two-dimensional echelle spectral data
was performed using the IRAF software package in a stan-
dard manner. Spectral data extracted from multiple ob-
ject images of V4641 Sgr were averaged in order to im-
prove the signal-to-noise (S/N) ratio. The wavelength
calibration was performed using the Th-Ar comparison
spectra obtained during the observations. The measured
FWHM of the weak Th lines was 0.15 A at 6000 A, and
the resulting resolving power was around R = 40000.
When the observed star shows broadened line profiles
due to its high rotational velocity, the process of contin-
uum fitting to the extracted raw spectral data has to be
carried out carefully. A very shallow and wide spectral
feature might be mistakenly interpreted as the continuum
level when a high-order polynomial function is employed
in the process. The task of continuum fitting the spec-
tral data for V4641 Sgr was carried out independently by
two of us (K. S. and M. T.-H.) using two different ap-
proaches. K. S. saved the fitted functions of each echelle
order obtained for the spectra of the sharp-lined stars 14
Cyg and ν Cap, and used them to divide the raw spec-
tral data of the corresponding echelle orders of V4641 Sgr.
Fitted functions of the continuum for 14 Cyg and ν Cap
were then applied to the V4641 Sgr observations obtained
on May 21 and June 20, respectively. M. T.-H. carefully
selected line-free windows in each echelle order for spec-
tra of the rapidly rotating spectrophotometric standard
star Feige 15 (Takada-Hidai, Aoki, and Zhao 2002), and
used them as reference to determine line-free windows in
echelle spectra of V4641 Sgr. Since Feige 15 is classified as
a spectral type of A0 in the SIMBAD database (which is
operated at CDS, Strasbourg, France), its effective tem-
perature can be regarded as being very close to that of
V4641 Sgr, with a spectral type of B9 III, suggesting a
spectroscopic similarity between the two stars. We found
good agreement between the results obtained from the two
methods, even for very weak features, and concluded that
we had correctly obtained normalized spectral data to be
used for the abundance analysis.
We compared the spectra of V4641 Sgr obtained on two
nights [May 21 and June 19, at photometric phases 0.38
and 0.62, respectively, which are calculated using data
given in Goranskij et al. (2003)] and found no detectable
spectral variation. Thus, the data observed on the two
nights were averaged in order to obtain the final spec-
trum, after correcting for the apparent doppler shifts due
to the orbital motion. The S/N ratios of the resulting
No. ]
Abundances in the Secondary Star of V4641 Sgr
3
spectrum of V4641 Sgr were measured at several contin-
uum windows near 5000 A and 6000 A. The averaged S/N
ratio (per pixel) were around 180 near 5000 A and 210
near 6000 A. For the two reference stars, a much higher
S/N ratio (around 400) was achieved at 6000 A.
3. Spectral Analysis
An abundance analysis of V4641 Sgr has been car-
ried out relative to the two reference stars 14 Cyg and
ν Cap. Table 1 lists the absorption lines used in the
following analyses. Log gf values are taken from the
VALD database (Kupka et al. 1999); atmospheric pa-
rameters (Teff and log g) for these two stars were taken
from Adelman et al. (2002). They are (Teff and log g)
= (10750 K, 3.5) and (10250 K, 4.0) for 14 Cyg and ν
Cap, respectively. Abundance analyses of these two stars
have been published in Adelman (1999) (14 Cyg) and in
Adelman (1991)(ν Cap); solar abundances are reported in
these two stars.
We estimated the effective temperature of V4641 Sgr
by comparing the strengths of the two He i lines at 4471
A and 5876 A in V4641 Sgr and in the two reference stars,
as shown in figure 1. In the figure, the observed spectra
of these three stars are compared with simulated ones.
Spectral simulations were performed using line-blanketed
model atmospheres interpolated from Kurucz (1993) and
assuming the solar abundances (except for figures 2 and 8,
in which an enhanced abundance of Mg by 0.20 dex is as-
sumed). In the spectral simulations, atomic data, except
for the major features listed in table 1, are taken from
Kurucz and Bell (1995). Equivalent widths of the two
He i lines were measured by directly integrating their pro-
files in the three stars. We found that equivalent widths
of both of the He i lines in V4641 Sgr are just around
the average of the two reference stars and conclude that
the Teff of V4641 Sgr is near the average of the two stars
(10500 ± 200 K). This is in agreement with that obtained
in Orosz et al. (2001).
We adopt the surface gravity (log g = 3.5) of V4641 Sgr
given in Orosz et al. (2001), who obtained this value from
the measured widths of the Balmer lines. We use a micro-
turbulent velocity ξt = 1.6 km s−1 in ν Cap (Dworetsky
and Budaj 2000), and assume ξt = 2.0 km s−1 in 14 Cyg.
It is difficult to determine the microturbulent velocity in
V4641 Sgr because we cannot measure the weak metallic
lines, which are in any case unaffected by a change in ξt,
in this star. The microturbulent velocity was guessed to
be around 2.5 ± 0.5 km s−1 by simulating the spectral re-
gion between 4505 and 4535 A , where moderately strong
(80 - 150 mA) Fe ii lines can be found, to obtain the best
fitting, assuming an appropriate rotational velocity (v sin
i, described in the next section).
4. Results
We estimate the rotational velocity, v sin i, of V4641
Sgr from the observed profile of the Mg ii line at 4481
A , as shown in figure 2. First, we tried to reproduce the
profiles of the line in 14 Cyg and ν Cap by adopting the
mean values of published data of v sin i. They are 30
and 25 km s−1 for 14 Cyg (Abt and Morrell 1995; Royer
et al. 2002) and ν Cap (Abt et al. 2002; Dworetsky and
Budaj 2000; Fekel 2003; Royer et al. 2002), respectively.
We find that acceptable reproductions of the observed line
profile can be obtained using adopted values of v sin i for
the two stars, as shown in figure 8. Next, we searched
for a suitable value of v sin i in V4641 Sgr by exploring
values between 60 and 150 km s−1; we found that the
best fit is achieved at 95 ± 10 km s−1. If we use v sin i
= 123 km s−1, as obtained in Orosz et al.
(2001), the
simulated profile becomes too shallow and too wide (figure
2). We conclude that their result of v sin i is too large,
most probably resulting from the low resolution of their
spectral data.
We tried to estimate the abundance of C in V4641 Sgr
using the C ii line at 4267 A (figure 3). The line consists
of two major components at 4267.001 A and at 4267.261
A (table 2). The C ii feature in ν Cap can be reproduced
with a solar abundance of C, while the observed feature in
14 Cyg appears to be too weak. This may suggest either
a slight under -- abundance of C or an error in the adopted
atmospheric parameters for 14 Cyg. We find that the
broard feature at 4267 A in V4641 Sgr can be reproduced
by assuming a solar abundance of C.
We find a shallow and broard absorption feature at
6483 A in V4641 Sgr in both the May 20 and June
19 data. Comparisons with the two reference stars (fig-
ure 4) shows that the feature in V4641 Sgr appears to
be too deep. The feature coincides with the positions of
the five components of N i lines (Multiplet No. 21). We
simulated spectra of both reference stars, assuming solar
abundances, and find that observations can be reasonably
reproduced when excluding the sharp components due to
atmospheric absorption. The only observable feature is
the Fe ii line at 6482.204 A. On the other hand, the ob-
served feature of V4641 Sgr at 6483 A cannot be repro-
duced when we assume solar abundances of N and Fe. A
reasonable reproduction can be obtained when a signifi-
cant over -- abundance of N (by around 1.0 dex) is assumed,
as shown in figure 4. Allowing for a slight uncertainty in
the continuum level because of the low S/N ratio, we con-
clude that N is indeed over -- abundant in V4641 Sgr by
at least 0.8 dex. This conclusion is in accordance with
the result noted in Orosz et al.
(2001), who obtained an
over -- abundance of N by 1.0 dex from an analysis of the
same spectral feature.
Analyses of weak O i lines near 6455 A are shown
in figure 5. We conclude that the O i feature can be
reasonably fit by using a solar O abundance. This is in
contrast with the result given in Orosz et al. (2001), who
obtained an over -- abundance of O (by a factor of three),
when analysing the same spectral feature.
Next, we analyse the Na i D lines. The D lines are usu-
ally contaminated by interstellar absorption superposed
on the stellar components, as shown for 14 Cyg and ν
Cap (figure 6). Fortunately, spectral lines of V4641 Sgr
observed on July 19 are significantly redshifted (by +210
4
K. Sadakane et al.
[Vol. ,
km s−1), and we can analyse both of the D1 and D2 com-
ponents, which are unaffected by interstellar absorptions.
The broad D lines in V4641 Sgr are too strong to be ac-
counted for by the solar Na abundance. As is shown in
figure 6, we need to enhance the abundance of Na by 0.7
dex in order to fit the D2 line. Further enhancement (by
0.2 dex) of the Na abundance is needed to fit the D1 line
(although the D1 line is heavily affected by atmospheric
absorption). Upon averaging the results obtained from the
D1 and D2 lines, we concluded that Na is over -- abundant
by at least 0.8 dex in V4641 Sgr.
The abundances of six elements (Mg, Si, Al, Ti, Cr, and
Fe) in V4641 Sgr were estimated by comparing the major
absorption features of each element with those seen in the
two reference stars. figures 7 to 11 display the results. In
these figures, the solid lines show simulated spectra for
each star, assuming solar abundances (except for figure 8,
where an enhanced abundance of Mg is assumed), taken
from Grevesse and Sauval (1998). Neutral and singly ion-
ized lines of Mg are shown in figures 7 and 8, respectively.
We find that the Mg i triplet lines can be reproduced with
a solar abundance of Mg, while a slight enhancement (0.20
dex) of Mg is suggested from the observed profile of the
Mg ii line at 4481 A. On the other hand, the Mg ii feature
at 4390.5 A in V4641 Sgr can be reproduced by assum-
ing a solar Mg abundance. Thus, we conclude that the
abundance of Mg in V4641 Sgr coincides with that of the
Sun, which differs from the results reported by Orosz et al.
(2001), who concluded that this star exhibits an enhanced
(by a factor of seven) abundance of Mg.
We find that an Al ii line at 4663 05 A is clearly present
in both of the two reference stars, and in V4641 Sgr (figure
9). Our spectral analyses of this line suggests that Al
may by slightly over -- abundant (by +0.2 dex) in V4641
Sgr, although the relatively poor S/N ratio in the blue
region does not allow for a reliable analysis of the line.
A pair of Si ii lines near 5050 A are compared in figure
10. We conclude that the abundance of Si in V4641 Sgr
is close to the solar abundance. In figure 11, we compare
Ti ii, Cr ii, and Fe ii
lines, where all of the absorption
features can be reproduced by assuming the same (solar)
abundances in the three stars. We analyse several other
strong Ti ii, Cr ii, and Fe ii
lines (noted in table 1)
and found that all of these features in V4641 Sgr can be
explained using solar abundances of Ti, Cr, and Fe. Our
result for Ti is again in contrast with the result given in
Orosz et al.
(2001), who obtained an over -- abundance
of Ti (by a factor of 10). They used four Ti ii lines (at
5129.2 A, 5185.9 A, 5188.7 A, and at 5226.5 A) to obtain
the Ti abundance, and pointed out a high abundance of
Ti from the two lines (at 5129.2 A and 5226.5 A) using
data of spectral resolution of about 4 A. We examined all
of these lines on our high-resolution data and find that all
these lines are very weak in the reference stars, and also
in V4641 Sgr, when compared to those Ti ii lines listed in
table 1. We concluded that the four lines used in Orosz et
al.
(2001) are inadequate to be used for the abundance
analysis of Ti.
Our final derived abundances for 10 observed elements
(11 ions) are summarized in table 2, together with their
expected errors. We estimate uncertainties in the abun-
dances of each ion introduced by errors in the adopted
parameters: 200 K in Teff, 0.5 in log g, and 0.5 km s−1
in ξt. When these errors are combined, we conclude that
the our abundance analysis results are reliable within 0.25
dex (table 2). We examined the effect of a difference in
spectral resolution on the resulting abundances by a sim-
ple test. The original data of both 14 Cyg and V4641 Sgr
were degraded to around R = 8000, the highest resolution
used in Orosz et al. (2001), by convolving with an appro-
priate Gaussian function. We then repeated abundance
analyses using the degraded data for several spectral fea-
tures such as the Mg ii line at 4481 A , the Mg i triplet
lines, and the three Ti ii lines listed in table 1. Fairly
good agreements (within 0.05 dex) were found for 14 Cyg
from both high and low resolution data. On the other
hand, differences as large as 0.15 dex were found between
abundance results obtained from weak and noisy spectral
features in the case of V4641 Sgr. We infer that these
differences are mainly resulted from the relatively poor
SN ratio in the V4641 Sgr data, but not from the differ-
ence in the spectral resolution. When the limited S/N
ratio of our observation and the high rotational velocity
of V4641 Sgr are taken into account, the expected error
in the abundances should be increased to around 0.3 dex.
5. Discussion
We obtained abundances of 10 elements in V4641 Sgr,
and found definite over -- abundances of only two elements
(N and Na). The abundances of the eight other elements
in V4641 Sgr have been shown to be the same as those
in the two reference stars (solar abundances), except for a
possible enhancement of Mg suggested from the Mg ii line
at 4481 A and that of Al. However, when averaged with
the result obtained from the Mg i triplet lines and the
Mg ii line at 4390.5 A, the abundance of Mg is coincident
with that in the reference stars within the expected error
([Mg/H] = +0.10 ± 0.30). The above conclusions are in
contrast to the results noted in Orosz et al.
(2001) ex-
cept for N, where they concluded over -- abundances of N
(1.0 dex), O (0.48 dex), Mg (0.85 dex), and Ti (1.0 dex)
in V4641 Sgr when compared to the Sun. We suggest that
the primary reason for obtaining discordant abundances
for O, Mg, and Ti is the difference in the spectral res-
olution of the data. Orosz et al.
(2001) used a much
lower resolving power (R ranging from 1200 to 7700) than
obtained in the present study (R ∼ 40000).
Our results for the abundances of the light elements in
the secondary star of V4641 Sgr [definite over -- abundances
of N and Na, normal (solar) abundances of O, and the α-
elements Mg, Si, and Ti] are unique when compared with
the results obtained for other X-ray binaries. Abundances
obtained in four secondary stars in X-ray binaries are com-
pared in figure 12 [V4641 Sgr (this study), GRO J1655-40
(Israelian et al. 1999), A0620-00 (Gonz´alez Hern´andez et
al. 2004), and Cen X4 (Gonz´alez Hern´andez et al. 2005)].
We note that all four stars show distinct abundance pat-
No. ]
Abundances in the Secondary Star of V4641 Sgr
5
terns. The α-elements (O, Mg, Si, S, and Ti) are def-
initely over -- abundant in GRO J1655-40, while they ap-
pear to be normal in V4641 Sgr. Fe is over -- abundant
only in Cen X-4. N and Na are over -- abundant only in
V4641 Sgr. The difference in the abundances of O be-
tween V4641 Sgr and GRO J1655-40 is impressive. The
observed abundance pattern in V4641 Sgr (enhanced N
and Na, and normal α-elements) seems to be different
from those of the usual supernova models that predict the
enhancement of α-elements (Podsiadlowski et al.
2002,
Gonz´alez Hern´andez et al. 2004, and Gonz´alez Hern´andez
et al. 2005). However, these variations of abundance pat-
ters can be explained with the variations of the abun-
dances of supernova explosions that are associated with
black hole formation (Umeda and Nomoto 2003).
In order to explain the abundances in V4641 Sgr quan-
titatively, we calculated the evolution of the star with
the initial mass of 40 M⊙ and the solar metallicity from
the main-sequence to collapse as in Umeda and Nomoto
(2005) and constructed the following evolutionary models
for V4641 Sgr.
In the close binary system, this 40 M⊙
primary star underwent a common envelope phase and
lost most of its H-rich envelope until it became a He star
of mass 15.14 M⊙. The system also lost its angular mo-
mentum and became compact with the orbital period as
short as that observed. Figure 13 shows the abundance
distribution near the surface of the He star at the on-
set of collapse. In the He-rich layer, 14N and 23Na were
enhanced by the CNO-cycle and Ne-Na cycle (proton cap-
tures on 21Ne and 22Ne) during H-burning. In the deeper
He layer, the 14N abundance was decreased by successive
α-capture to produce 22Ne during weak He shell burning.
The 15 M⊙ He star is massive enough to eventually
formed a black hole (BH). We assume that the collapse
induced a relatively weak explosion. Generally, an explo-
sion with a smaller energy leads to a larger amount of fall
back materials and thus a smaller amount of ejecta (e.g.,
Iwamoto et al. 2005). In order to reproduce the observa-
tions of V4641 Sgr, we assume that the kinetic energy of
explosion was as small as E = 6 ×1049 ergs. In such a weak
explosion, only 0.5 M⊙ materials above Mr = 14.66M⊙
were ejected. The abundance distribution in the He layer
in figure 13 does not change in the explosion. A part of
the ejected materials must be captured by the secondary
star. The captured (accreted) materials were then mixed
with the materials of solar metallicity in the atmosphere
of the secondary star. Since the ejecta is relatively N-
and Na-rich without any enhancement of α-elements, this
could explain the observed abundance pattern of V4641
Sgr. If the accreted material is mixed with 40-times larger
amount of secondary star materials, the final abundance
pattern would be consistent with the observed abundance
pattern of the secondary star (figure 14). Here, most of the
heavy elements above Na originated from the materials of
V1641 Sgr. Such a partial mixing (e.g., slow rotational
mixing) may be realized because the surface temperature
of V4641 Sgr is too high for deep convective mixing to
occur.
We should note that an alternative scenario is possi-
ble. If the stellar wind of the He star of ∼ 15 M⊙ blows
at a high enough rate, a part of the N- and Na-rich ma-
terials in the He layer would have been blown off and
captured by the secondary star. If the energy of super-
nova explosion was even smaller than ∼ 6 × 1049, no mass
ejection occurred. These processes could lead to the ob-
served abundance pattern of V4641 Sgr. The above 40
M⊙ model formed a BH of 14 M⊙. The initially 30 M⊙
model produces a similar abundance pattern by forming
a 7.2 M⊙ BH. Since the observed BH mass of V4641 Sgr
is ∼ 9.6M⊙, the progenitor of the BH in V4641 is likely a
∼ 35M⊙ star. It is highly uncertain in the current super-
nova models under what condition the BH formation can
induce a supernova explosion and how much explosion en-
ergy can be released; it may depend on the rotation of the
BH and the progenitor. The case of V4641 Sgr suggests
the BH- forming supernova was really dark, because no
radioactive 56Ni was ejected. Such a dark supernova cor-
responds to the extreme end of the faint supernova branch
(Nomoto et al. 2005).
Another possible scenario is the contamination by ro-
tationally induced mixing in the secondary star, itself.
However, the observed rotational velocity (v sin i ∼ 100 km
s−1) and estimated mass may be lower than those pre-
dicted for the simultaneous enhancements of N and Na,
although they strongly depend on the uncertain inclina-
tion.
We thank Drs. K. Matsumoto and M. Uemura for com-
ments and suggestions. This research was partly sup-
ported by Grants-in-Aid from the Ministry of Education,
Culture, Sports, Science and Technology (No. 15540236
KS, No.
17540218 MTH, No. 17740163 NI, and No.
17033002 KN). T.C.B. ackowledges partial support from
grant AST 04-06784, as well as from grant PHY 02-
16783, Physics Frontier Center/Joint Institute for Nuclear
Astrophysics (JINA), awarded by the US National Science
Foundation.
References
Abt, H. A., Levato, H., & Grosso, M. 2002, ApJ, 573, 359
Abt, H. A., & Morrell, N. I. 1995, ApJS, 99, 135
Adelman, S. J. 1991, MNRAS, 252, 116
Adelman, S. J. 1999, MNRAS, 310, 146
Adelman, S. J., Pintado, O. I., Nieva, F., Rayle, K. E., &
Sanders, S. E., Jr. 2002, A&A, 392, 1031
Dworetsky, M. M., & Budaj, J. 2000, MNRAS, 318, 1264
Fekel, F. C. 2003, PASP, 115, 807
Gonz´alez Hern´andez, J. I., Rebolo, R., Israelian G., & Casares,
J. 2004, ApJ, 609, 988
Gonz´alez Hern´andez, J. I., Rebolo, R., Israelian G., Casares,
J., Maeda, K., Bonifacio, P., & Molaro, P. 2005, ApJ, 630,
495
Goranskij, V. P., Barsukova, E. A., & Burenkov, A. N. 2003,
Astronomy Rept., 47, 740
Grevesse, N., & Sauval, A. J. 1998, Space Sci. Rev., 85, 161
Israelian, G., Rebolo, R., Basri, G., Casares, J., & Martin, E.
L. 1999, Nature, 401, 142
6
K. Sadakane et al.
[Vol. ,
Iwamoto, N., Umeda, H., Tominaga, N., Nomoto, K., &
Maeda, K. 2005, Science, 309, 451
Kato, T., Uemura, M., Stubbings, R., Watanabe, T., &
Monard, B. 1999, Inf. Bull. Variable Stars, 4777
Kupka, F., Piskunov, N., Ryabchikova, T. A., Stempels, H.C.,
& Weiss, W.W. 1999, A&AS, 138, 119
Kurucz, R. L.
1993, Kurucz CD-ROM, No.13 (Harvard-
Smithsonian Center for Astrophysics)
Kurucz, R. L., & Bell, B.
1995, Kurucz CD-ROM, No.23
(Harvard-Smithsonian Center for Astrophysics)
Noguchi, K., et al. 2002, PASJ, 54, 855
Nomoto, K., Tominaga, N., Umeda, H., Maeda, K., Ohkubo,
T., & Deng, J. 2005, Nuclear Phys., A 758, 263c
Orosz, J. A., et al. 2001, ApJ, 555, 489
Podsiadlowski, P., Nomoto, K, Maeda, K., Nakamura, T.,
Mazzali, P., & Schmidt, B. 2002, ApJ, 567, 491
Royer, F., Grenier. S., Baylac, M. -O., Gomez, A. E., & Zorec,
J. 2002, A&A, 393, 897
Smith, D. A., Levine, A. M., & Morgan, E. H., 1999, IAU
Circ., 7253, 2
Takada-Hidai, M., Aoki, W., & Zhao, G. 2002, PASJ, 54, 899
Uemura, M., et al. 2002a, PASJ, 54, 95
Uemura, M., et al. 2002b, PASJ, 54, L79
Uemura, M., et al. 2004a, PASJ, 56, S61
Uemura, M., et al. 2004b, PASJ, 56, 823
Uemura, M., et al. 2005, Inf. Bull. Variable Stars, 5626
Umeda, H., & Nomoto, K. 2003, Nature, 422, 871
Umeda, H., & Nomoto, K. 2005, ApJ, 619, 427
No. ]
Abundances in the Secondary Star of V4641 Sgr
7
Table 1. Lines used in abundance analyses.
He i
C ii
N i
O i
Na i
Mg i
Mg ii
Al ii
Si ii
Ti ii
Cr ii
Fe ii
Wavelength(A)
4471.469
4471.473
4471.473
4471.485
4471.488
4471.682
5875.599
5875.614
5875.615
5875.625
5875.640
5875.966
4267.001
4267.261
4267.261
4267.261
6480.512
6481.706
6482.699
6483.753
6484.808
6453.602
6454.444
6455.977
5889.951
5895.924
5167.321
5172.684
5183.604
4390.514
4390.572
4481.126
4481.325
4663.046
5041.024
5055.984
5056.317
4395.000
4501.270
4571.960
4558.660
4588.220
4592.070
4508.280
4515.340
4576.330
4582.840
4583.830
4629.340
χ(eV)
20.964
20.964
20.964
20.964
20.964
20.964
20.964
20.964
20.964
20.964
20.964
20.964
18.046
18.046
18.046
18.046
11.753
11.750
11.764
11.753
11.758
10.740
10.740
10.741
0.00
0.00
2.709
2.712
2.717
9.99
9.99
8.864
8.864
10.598
10.067
10.074
10.074
1.084
1.116
1.572
4.073
4.071
4.074
2.856
2.844
2.844
2.844
2.807
2.807
log gf
-2.198
-1.028
-0.278
-1.028
-0.548
-0.898
-1.511
-0.341
0.409
-0.341
0.139
-0.211
0.562
-0.584
0.717
-0.584
-1.420
-1.063
-0.510
-0.857
-0.674
-1.288
-1.066
-0.920
0.117
-0.184
-1.030
-0.402
-0.180
-1.490
-0.530
0.740
0.590
-0.284
0.291
0.593
-0.359
-0.510
-0.760
-0.230
-0.449
-0.627
-1.221
-2.250
-2.450
-2.920
-3.090
-1.860
-2.330
Table 2. Abundances in V4641 Sgr.
[X/H]
Error(A)
Error(B)
Error(C)
N i
C ii
0.0 +0.8
0.02
0.12
0.0
0.27
0.01
0.02
O i Na i Mg i Mg ii Al ii
+0.1 +0.2
0.0 +0.8
0.04
0.03
0.12
0.01
0.14
0.19
0.16
0.01
0.0
0.13
0.02
0.08
0.0
0.13
0.18
0.10
Note. -- [X/H]: logarithmic abundances relative to the
Sun; Error (A), (B), and (C): expected uncertainties in
the abundance introduced by errors in Teff (200 K), log g
(0.5), and in ξ t (0.5 km s−1), respectively.
4466
!"#
$ %&'
4468
4470
4472
4474
4476
5872
5874
5876
5878
5880
()*+,-./01 234
Fig. 1. Comparisons of He i lines. Profiles of two He i lines at
4471 A and at 5876 A in V4641 Sgr and in two comparison
stars 14 Cyg and in ν Cap are compared in the upper and
lower panels, respectively. The dots and solid lines show the
observation and the simulated spectra calculated assuming
the solar abundances, respectively. Rotational velocities, v
sin i, of 30, 25, and 95 km s−1 are used for 14 Cyg, ν Cap,
and V 4641 Sgr, respectively.
Si ii Ti ii Cr ii Fe ii
0.0
0.0
0.04
0.03
0.13
0.06
0.05
0.10
0.0
0.10
0.03
0.13
0.0
0.05
0.10
0.09
5
6
7
8
9
:
;
<
=
>
?
@
A
B
C
D
E
F
8
K. Sadakane et al.
[Vol. ,
4478
4482
4480
GHIJKLMNOP QRS
4484
xy z{
4260
]^ _`a
bcdef ghi
j klm
n op
tu vw
qr s
4265
4270
}~
4275
Fig. 2. Rotational velocity, v sin i, of V4641 Sgr. A small
section of the observed spectrum near the Mg ii line at 4481.2
A is shown by dots. Simulated spectra, calculated assuming
the abundance of Mg, [Mg/H] = 0.2, for v sin i = 130, 95,
and 60 km s−1 are shown by dashed, thick solid, and thin
solid lines, respectively.
Fig. 3. Analysis of the C ii feature at 4267.2 A. The dots
and thin solid lines show the observation and the simulated
spectra calculated assuming the solar abundances of C. The
thick dashed line (for V4641 Sgr) show the case when C is
over -- abundant by 0.7 dex.
T
U
V
W
X
Y
Z
[
\
No. ]
Abundances in the Secondary Star of V4641 Sgr
9
¡¢
ýþ ÿ
õö÷øù úûü
6470
6475
6480
6485
6490
6495
£¤¥¦§©ª«¬ -®¯
6450
6452
6454
6456
ßàáâãäåaeçè éêë
6458
6460
Fig. 4. Analysis of the N i feature near 6483 A. The dots
and thin solid lines show the observation and the simulated
spectra calculated assuming the solar abundances of N. Many
sharp absorption features are caused by atmospheric H2O
molecules. The thick solid lines and the dashed line (for
V4641 Sgr) show the case when N is over -- abundant by 1.0
dex and 1.5 dex, respectively.
Fig. 5. Analysis of O i lines near 6455 A. The dots and solid
lines show the observation and the simulated spectra calcu-
lated assuming solar abundances of O and Fe, respectively.
The dashed line for V4641 Sgr is the case when O is 0.5 dex
over -- abundant as noted in Orosz et al. (2001).
°
±
²
³
´
µ
¶
·
¸
¹
º
»
¼
½
¾
¿
À
Á
Â
Ã
Ä
Å
AE
Ç
È
É
Ê
Ë
Ì
Í
Î
Ï
Ð
Ñ
Ò
Ó
Ô
Õ
Ö
×
Ø
Ù
Ú
Û
Ü
Ý
Þ
ì
í
î
ï
ð
ñ
ò
ó
ô
10
K. Sadakane et al.
[Vol. ,
5678 9:;< =>
-./0
1234
5880
5885
5890
5895
5900
!"#
LM NOP
QRSTU VWX
Y Z[\
BC D
?@ A
EF G
HI JK
5165
5170
5175
5180
5185
]^_`abcdef ghi
Fig. 6. Analysis of the Na i D lines. Both D1 and D2 lines
in 14 Cyg and ν Cap are contaminated by sharp interstellar
(I.S.) components. The I.S. lines of V4641 Sgr observed on
2005 June 19 are apparently blue-shifted by 210 km s−1 due to
the orbital motion of the star. The dots and thin solid lines
show the observation and the simulated spectra calculated
assuming the solar abundances of Na. The dashed line for
V4641 Sgr shows the case when Na is over -- abundant by 0.7
dex.
Fig. 7. Analysis of the Mg i triplet lines. Dots and solid lines
show the observation and the simulated spectra calculated
assuming solar abundances of Mg and Fe, respectively. The
dashed line for V4641 Sgr is the case when Mg is 0.85 dex
over -- abundant as noted in Orosz et al. (2001).
$
%
&
'
(
)
*
+
,
j
k
l
m
n
o
p
q
r
No. ]
Abundances in the Secondary Star of V4641 Sgr
11
st uvw
}~
x yz{
4476 4478 4480 4482 4484 4486
¡¢
£¤¥¦§ ©ª
« ¬-®
·¸ ¹º
¯° ±²
³´ µ¶
4655
4660
4665
4670
»¼½¾¿ÀÁÂÃÄ ÅAEÇ
Fig. 8. Analysis of the Mg ii line at 4481 A. The dots rep-
resent observed spectra. The solid lines for 14 Cyg and ν
Cap are simulated spectra calculated assuming the solar abun-
dance of Mg, while that for V4641 Sgr is calculated assuming
an enhanced (by 0.20 dex) abundance of Mg. The dashed line
for V4641 Sgr is the case when Mg is 0.85 dex over -- abundant
as noted in Orosz et al.
(2001).
Fig. 9. Analysis of the Al ii line at 4663 A. The dots and
solid lines show the observation and the simulated spectra
calculated assuming a solar abundances for Al, respectively.
The dashed line for V4641 Sgr is the case when Al is 0.5 dex
over -- abundant.
È
É
Ê
Ë
Ì
Í
Î
Ï
Ð
12
K. Sadakane et al.
[Vol. ,
ÙÚ ÛÜÝ
Þssàáâ ãäå
ae çèé
ÑÒ ÓÔ
!" #$ %& '(
5035
5040
5045
5050
5055
5060
4570 4575 4580 4585 4590 4595 4600
êëìíîïðñòó ôõö
)*+,-./012 345
ÕÖ ×Ø
Fig. 10. Analysis of a pair of Si ii lines near 5050 A. The dots
and solid lines show the observation and the simulated spectra
calculated assuming the solar abundance of Si, respectively.
The dashed line for V4641 Sgr is the case when Si is 0.5 dex
over -- abundant.
Fig. 11. Analysis of Ti ii, Cr ii, and Fe ii lines near 4585
A. The dots and solid lines show the observation and the
simulated spectra calculated assuming solar abundances of
Ti, Cr, and Fe, respectively. The dashed line for V4641 Sgr
is the case when Ti, Cr, and Fe are 0.2 dex over -- abundant.
÷
ø
ù
ú
û
ü
ý
þ
ÿ
6
7
8
9
:
;
<
=
>
No. ]
Abundances in the Secondary Star of V4641 Sgr
13
4
P
8
,
$
$
,
% 7 0
12
16
20
DEFGHI JKLMNO
24
28
Fig. 13. Internal abundance distribution of the solar metal-
licity 15.14M⊙ He star model at the onset of collapse. This
He star is the core of the 40M⊙ star. The extent of the large
scale fallback is Mr = 14.66M⊙ shown by right-headed arrow.
The remaining black hole mass is 14.66M⊙. The main con-
tribution for Na as well as for N comes from H-burning. N
and Na abundances in the He layers are enhanced with re-
spect to the initial abundances of 1.2 × 10−3 and 3.5 × 10−5,
respectively.
Fig. 12. Comparison of metal abundances.
[X/H] values,
logarithmic abundance of the element X relative to the Sun,
in four secondary stars are plotted against the atomic number.
Filled circles, open circles, open triangles, and open squares
are for V4641 Sgr, GRO J1655-40, A0620-00, and for Cen
X-4, respectively.
Fig. 14. Comparison of V4641 Sgr with a mixture of the
solar metallicity 15.14M⊙ He star supernova ejecta and the
secondary star materials. The assumed ratio of the super-
nova ejecta to the secondary star materials is 1/40. The
filled circles with error-bars show the observed abundances.
The abundance pattern of the mixture is represented by open
squares connected with solid lines. The yield shows that, ex-
cept for the enhancements of N and Na, all the other elements
have almost solar ratios. Black hole with the mass of 14.66M⊙
is left behind.
?
@
A
B
C
|
0710.3117 | 1 | 0710 | 2007-10-16T16:35:14 | Applicability of colour index calibrations to T Tauri stars | [
"astro-ph"
] | We examine the applicability of effective temperature scales of several broad band colours to T Tauri stars (TTS). We take into account different colour systems as well as stellar parameters like metallicity and surface gravity which influence the conversion from colour indices or spectral type to effective temperature. For a large sample of TTS, we derive temperatures from broad band colour indices and check if they are consistent in a statistical sense with temperatures inferred from spectral types. There are some scales (for V-H, V-K, I-J, J-H, and J-K) which indeed predict the same temperatures as the spectral types and therefore can be at least used to confirm effective temperatures. Furthermore, we examine whether TTS with dynamically derived masses can be used for a test of evolutionary models and effective temperature calibrations. We compare the observed parameters of the eclipsing T Tauri binary V1642 Ori A to the predictions of evolutionary models in both the H-R and the Kiel diagram using temperatures derived with several colour index scales. We check whether the evolutionary models and the colour index scales are consistent with coevality and the dynamical masses of the binary components. It turns out that the Kiel diagram offers a stricter test than the H-R diagram. Only the evolutionary models of Baraffe et al. (1998) with mixing length parameter 1.9 and of D'Antona & Mazzitelli (1994, 1997) show consistent results in the Kiel diagram in combination with some conversion scales of Houdashelt et al. (2000) and of Kenyon & Hartmann (1995). | astro-ph | astro-ph | Astron. Nachr./AN xxx (xxxx) x, xxx -- xxx
7
0
0
2
t
c
O
6
1
]
h
p
-
o
r
t
s
a
[
1
v
7
1
1
3
.
0
1
7
0
:
v
i
X
r
a
Applicability of colour index calibrations to T Tauri stars
TORSTEN SCH ONING1 and MATTHIAS AMMLER1
,
2
1 Astrophysical Institute and University Observatory Jena, Schillergasschen 2-3, 07745 Jena, Germany
2 Centro de Astronomia e Astrof´ısica da Universidade de Lisboa, Tapada da Ajuda, 1349-018 Lisboa, Portugal
Received <date>; accepted <date>; published online <date>
Abstract. We examine the applicability of effective temperature scales of several broad band colours to T Tauri stars (TTS).
We take into account different colour systems as well as stellar parameters like metallicity and surface gravity which influence
the conversion from colour indices or spectral type to effective temperature.
For a large sample of TTS, we derive temperatures from broad band colour indices and check if they are consistent in a
statistical sense with temperatures inferred from spectral types. There are some scales (for V − H, V − K, I − J, J − H,
and J − K) which indeed predict the same temperatures as the spectral types and therefore can be at least used to confirm
effective temperatures.
Furthermore, we examine whether TTS with dynamically derived masses can be used for a test of evolutionary models and
effective temperature calibrations. We compare the observed parameters of the eclipsing T Tauri binary V1642 Ori A to the
predictions of evolutionary models in both the H-R and the Kiel diagram using temperatures derived with several colour
index scales. We check whether the evolutionary models and the colour index scales are consistent with coevality and the
dynamical masses of the binary components. It turns out that the Kiel diagram offers a stricter test than the H-R diagram.
Only the evolutionary models of Baraffe et al. (1998) with mixing length parameter α = 1.9 and of D'Antona and Mazzitelli
(1994, 1997) show consistent results in the Kiel diagram in combination with some conversion scales of Houdashelt et al.
(2000) and of Kenyon and Hartmann (1995).
Key words: stars: pre-main sequence -- stars: late-type -- stars: fundamental parameters -- stars: statistics -- stars: individual
(V1642 Ori A)
c(cid:13)0000 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1. Introduction
Only in a few cases the effective temperature of T Tauri
stars (TTS) can be measured directly. In most other cases,
conversions from spectral type to effective temperature are
used (see for example Cohen and Kuhi 1979; Hartigan et al.
1994). However, the spectral type of TTS is not always well
determined due to their variability. For example, the classical
TTS DL Tau is classified as GVe -- K7Ve (Samus et al. 2004).
Thus, we recommend to verify the spectral type temperature
with at least one colour index conversion scale. In this pa-
per we want to carefully examine which colour indices are
most suited on variable young stars with possible UV and IR
excesses.
To illustrate the effects of the variability of TTS, two
examples shall be given. The classical TTS (cTTS) most
variable in V , DR Tau, has a variability in (B − V )J in
the range −0.76 ≤ (B − V )J ≤ 1.11 (for details, see
Correspondence to: [email protected]
Herbst et al. 1994). This difference corresponds roughly to
the difference between an O-type and a K-type star. As ex-
pected, the variability of weak-line TTS (wTTS) is much
smaller. The most variable wTTS, V410 Tau, has a variability
of 0.87 ≤ (B − V )J ≤ 1.32 (Herbst et al. 1994) which cor-
responds roughly to the difference between an early K-type
and a late K-type star. In contrast, the typical measurement
error of a colour index is only about 0.02 magnitudes.
The applied colour index scales are summarised in Sect.
2; the way we calculated the colour index temperatures is ex-
plained in Sect. 3.
In Sect. 4, a statistical test for a sample of apparently sin-
gle TTS is done by comparing their colour index tempera-
tures with their spectral type temperatures. We have not found
any similar test in the literature.
If, in addition to the temperature, another stellar para-
meter like log g or Lbol is known, one can estimate the mass
and the age of a star by means of evolutionary models. If fur-
thermore a mass determination independent from evolution-
2 DESCRIPTIONOFTHEAPPLIEDSCALES
2.2 Tabulatedscales
ary models is known, one can check whether predicted and
measured mass are consistent. For binaries, one can further-
more check whether both components are coeval. By using
various conversion scales, we test combinations of evolution-
ary model and conversion scale (for details, see Sect. 5).
2. Description of the applied scales
For our examination we used a compilation of common
colour scales. The aim of our work is not a complete study
of colour scales -- instead we are only interested which scales
are best suited to infer the temperature of TTS. We selected
only scales for broad band colour indices covering (at least)
most of the spectral range of TTS. For example, the scale of
Kenyon and Hartmann (1995) for V − N is excluded because
it only covers the range from A0 to G1. The main properties
of these scales are summarised in Table 1. In this table we
introduce abbreviations for the scales containing the abbrevi-
ated references and the used colour index. The last letter of
these abbreviations indicates whether it is a giant ("g"), inter-
mediate ("i"), or dwarf ("d") scale. In case of ambiguity due
to the use of multiple scales the abbreviations are extended
by an expression in brackets giving either the number or the
spectral range of the corresponding scale, e. g. AAMR99g(8).
In the literature, the conversion scales are given either as
polynomial scales or as tabulated scales. We did not only
adopt scales from the literature but in addition we defined
combined scales for such colour indices which are used for
at least two scales in order to be able to compare the different
colour indices directly. Annotations to the different types of
scales are given in the following subsections.
2.1. Polynomial scales
of
scales
di Benedetto
(1998),
The
Blackwell and Lynas-Gray (1994), Gratton et al.
(1996),
and Houdashelt et al. (2000) were derived by fitting one or
more polynomials TCI(X − Y )1 only to measured colour
indices and temperatures of individual stars.
the
in
use
polynomial
the metallicity as
order
Alonso et al. (1996, 1999a) and Sekiguchi and Fukugita
additional parameter
(2000)
to minimise
scatter.
for
Sekiguchi and Fukugita (2000)
is the only scale using
also the surface gravity log g as additional parameter. Thus
our own work has to account for the metallicity and the
surface gravity in order to infer the effective temperature
of a certain TTS. As the metallicity of TTS is usually
not known, we take the extreme values for stars of the
thin disc as measured by Fuhrmann (2004) as an estimate
([Fe/H] = −0.05 ± 0.55). If no measurements for the
surface gravity are available, i. e. for non-eclipsing binaries
and single stars, log g = 3 ± 2 is taken as rough estimate
containing all values predicted by evolutionary models for
TTS. For the scale of Sekiguchi and Fukugita (2000), this
quite large error yields a maximum error in the temperature
1 In this paper, X refers to any broad band magnitude and X − Y
refers to any broad band colour index
of only 170 K so that we can apply indifferently either giant
and dwarf scales to TTS.
The temperatures which are used to calibrate the scales
are usually derived by means of the InfraRed Flux Method
(IRFM, Alonso et al. 1999b), scaled to direct measurements
of the temperature. It is noteworthy that Gratton et al. (1996)
lowered the IRFM temperatures of Bell and Gustafsson
(1989) by 122 K to adjust
them to the temperatures of
Blackwell and Lynas-Gray (1994) while Bell and Gustafsson
(1989) themselves suggest a lowering by only 80 K.
The photometric data used to derive these scales
was taken from own measurements or
from the lit-
erature. Thereby the literature values were often con-
verted from other broad band filter systems. The K mag-
nitudes of Blackwell and Lynas-Gray (1994) were con-
verted from the narrow band Kn of Selby et al. (1988);
the B magnitudes used by Sekiguchi and Fukugita (2000)
were presumably converted from Stromgren photometry of
Hauck and Mermilliod (1998). Gratton et al. (1996) do not
mention the sources of their photometry.
The scales are based on different stellar samples. Between
13 and more than 500 stars were used.
Usually only a constant relative or absolute intrinsic
error for the whole relation is given. On the other hand,
Gratton et al. (1996) and Houdashelt et al. (2000) give errors
for the coefficients of the polynomial. This yields larger er-
rors (sometimes more than 1000 K) for cooler stars.
2.2. Tabulated scales
We consider two tabulated scales (Hartigan et al. 1994;
Kenyon and Hartmann 1995) giving temperatures, intrinsic
colours, and bolometric corrections compiled from the litera-
ture for different spectral types.
The temperatures in Kenyon and Hartmann (1995, ta-
ble A5) were taken from Schmidt-Kaler (1982). Most of the
given colours were either taken directly from various sources
or derived by interpolating linearly between those values. For
dwarfs from F- to M-type (which we are interested in) most of
the infrared colours were taken from Bessell and Brett (1988)
using the Johnson-Glass filter system, supplemented by data
of Johnson (1966). However, we cannot retrace how some
of the given colours were derived - especially the infrared
colours of late G type stars.
The temperatures in Hartigan et al. (1994, table 4) were
taken from Schmidt-Kaler (1982) for spectral types from F0
to K4 and from Bessell (1991) for K7 and integer M sub-
classes. For K5, M0.5, and M1.5 the origin of the tempera-
tures cannot be retraced. From this paper only the scales for
(R − I)C and IC − J are used, as no other scale could be
found for these colour indices. The values for (R − I)C for
integer M sub-classes were also taken from Bessell (1991),
the origin of the other values for (R − I)C as well as for all
IC − J cannot be retraced.
Hartigan et al. (1994) as well as Kenyon and Hartmann
(1995) do not give any intrinsic errors for these scales. We
adopt the errors derived in Ammler et al. (2005), namely
∆ log Teff = 0.015 for temperatures originally given by
Schmidt-Kaler (1982) and ∆Teff = 280 K for temperatures
originally given by Bessell (1991).
2.3. Combined scales
For these scales, the temperatures are the mean of the temper-
atures of all scales which yield a result for the corresponding
values of X − Y , log g, and [Fe/H]. The error of these com-
bined temperatures is composed of the mean error and the
standard deviation of the individual temperatures. These both
errors are considered independent of each other and thus are
added quadratically. These scales are hence labelled as "com-
bined X − Y ".
3. Remarks on the calculation of the colour
index temperatures TCI
In an ideal case, the colour index conversion scales should be
only used for observations done with the same filter system.
This is in general not possible because there are too many dif-
ferent filter systems. A consequent conversion into only one
standard filter system is not possible, because often either: (a)
The observational data are given in instrumental filter sys-
tems for which a conversion formula is not known. (b) The
photometry was already converted from another filter system.
With a further conversion one would have twice the statisti-
cal error of such a conversion. Or (c) the used filter system is
even not given.
Detailed information about
the different filter sys-
tems can be found in Moro and Munari
(2000) and
Fiorucci and Munari (2003)2; information about the TCS fil-
ter system can be found in Alonso et al. (1994, 1998).
Especially for the R and I band the differences between
the distinct filter systems are not negligible. Considering
V − R = 0.50, for example, the difference between the
Cousins and the Johnson filter system yields a difference of
about 1000 K in the resulting colour index temperature. Thus,
3 and (RI)J are treated as completely dif-
in this paper, (RI)C
ferent passbands. Observations in other RI filter systems are
ignored because corresponding colour index scales do not ex-
ist. All other passbands are always treated as if they were
Johnson, adding a negligible error.
The given colour indices have to be corrected for in-
terstellar reddening, i. e. they have to be converted to in-
trinsic colours. Therefore, one has to calculate the (filter
system dependent) extinction AX from the visual extinc-
tion AV which is usually given. For this purpose, the AX
AV
of Fiorucci and Munari (2003) for an interstellar reddening
RV = 3.1 are used. Possible systematic errors in the deter-
mination of the AV cannot be considered in this paper.
2 The filter
system "Bessell and Brett
Moro and Munari
equal to the Johnson-Glass filter system.
(2000) and Fiorucci and Munari
(1988)" given in
is
(2003)
3 For most TTS, R and I have been measured in the Cousins sys-
tem, probably due to the larger transmission of these filters.
4 THESTATISTICALTESTWITHSINGLETTS
4. The statistical test with single TTS
(1998). Thereby,
For a sample of about 150 apparently single TTS we calcu-
lated the difference D between the colour index temperature
TCI and the mean spectral type temperature TS for each star
and for each colour index conversion scale.
(1)
D = TCI − TS
TS is derived with the conversion scales of Bessell
(1979, 1991); Cohen and Kuhi
(1979); Schmidt-Kaler
(1982); de Jager and Nieuwenhuijzen (1987); Hartigan et al.
(1994); Kenyon and Hartmann (1995); Luhman (1999);
and Perrin et al.
the given spectral
types are converted into spectral numbers (following
de Jager and Nieuwenhuijzen 1987). For example, M1 is
converted to 6.7. If necessary, we interpolated linearly, with
spectral types of the conversion scales being converted into
spectral numbers. If intervals are given, the mean spectral
number is used. For example, instead of "K7-M0" we use
the spectral number 6.35, which corresponds to "K8.5". The
error of the spectral number of our sample stars is either
assumed to be 0.05 (according to roughly half a subclass)
or as half the given interval. The TS uncertainty ∆TS is
composed of the mean intrinsic error of the used scales (see
Ammler et al. 2005, table 2, for details) and the standard
deviation of the individual spectral type temperatures.
We then used four criteria to test whether a scale is appli-
cable to TTS or not (see Sect. 4.1 for details).
For comparison, we did the same test with a sample of
about 250 non-variable main sequence and / or giant stars
(hence referred to as "old stars"). Of course, in this case dwarf
scales were only used for dwarfs and giant scales only for
giants.
The data for our sample of single TTS (photometry,
AV , and spectral type) were taken from a compilation of
Ralph Neuhauser (priv. comm.)4 We have taken the data for
the old stars from the "Catalogue of the Brightest Stars"
of Ochsenbein and Halbwachs (1987), whereas only non-
variable stars with known V magnitude, spectral type F0 or
later, and luminosity class III to V were used. The (UBVRI)J
photometry of this catalogue was supplemented with 2MASS
JHK photometry. The visual extinction AV of the old stars
was calculated from the Hipparcos parallax and the distance
dependent extinction γ = 0.3 mag kpc−1 (valid in the galac-
tic plane outside of dark clouds). For both samples, 0.02 mag-
nitudes are adopted as typical measurement error of the re-
sulting intrinsic colours.
(1993,
4 which
comprises
data
(1992), Briceno et al.
(1979), Elias
(1992), Haas et al.
(1994), Hartmann et al.
from Basri and Marcy
1998,
(1978), Gizis et al.
(1990), Hartigan
(1994),
Beichman et al.
1999),
Cohen and Kuhi
(1999),
Gomez et al.
(1993),
Hartigan et al.
(1991), Herbig and Bell
(1988), Herbig et al. (1986), Jones and Herbig (1979), Kenyon et al.
(1990, 1993, 1994a,b), Leinert and Haas (1989), Leinert et al.
(1994), Martin et al.
(1991,
(1987),
(1994), Moneti and Zinnecker
Reid and Hawley (1999), Rydgren (1984), Simon et al.
(1992,
1995), Skrutskie et al. (1990), Strom et al.
(1989), Strom et al.
(1990), Strom and Strom (1994), Torres et al. (1995), Vrba et al.
(1985), and Walter et al. (1988).
1993), Magazzu and Martin
(1991), Myers et al.
4
T
H
E
S
T
A
T
I
S
T
I
C
A
L
T
E
S
T
W
I
T
H
S
I
N
G
L
E
T
T
S
Table 1. Principal characteristics of the scales taken from the literature. Further details see Sect. 2.
Reference
Abbreviation(s)
Alonso et al. (1996)
Alonso et al. (1999a)
di Benedetto (1998)
Blackwell and Lynas-Gray (1994)
Gratton et al. (1996)
Hartigan et al. (1994)
Houdashelt et al. (2000)
Kenyon and Hartmann (1995)
AAMR96d
AAMR99g
Ben98d
Ben98i
Ben98g
BLG94i
GCC96d
GCC96g
HSS94d
HBS00d
HBS00g
KH95d
Sekiguchi and Fukugita (2000)
SF00i
1 -- Errors are taken from Ammler et al. (2005)
Colour Indices
Considered In Our Work
(B − V ), (V − R)J, (V − I)J, (V − K),
(R − I)J, (J − H)TCS, (J − K)TCS
(U − V )J, (B − V )J, (V − R)J, (V − I)J,
(V − K)TCS, (V − L′)TCS, (R − I)J, (I − K)J,
(J − H)TCS, (J − K)TCS,
(V − K)J
(V − K)J
(B − V )J, (V − R)J, (R − I)J, (J − K)J,
(V − K)J
(R − I)C, IC − J
(U − V )J, (B − V )J, (V − R)C, (V − I)C,
(V − K)CIT, (V − K)JG, (J − K)CIT, (J − K)JG
(U − V ), (B − V ), (V − R)C, (V − R)J,
(V − I)C, (V − I)J, (V − J)JG, (V − H)JG,
(V − K)JG, (V − L)JG, (V − M)JG
(B − V )
Considered Range
Of Spectral Types
F0 -- K5
F0 -- K5
F -- K
A -- M
F -- K
F0 -- M6
F -- K
F0 -- M6
F0 -- K5
Type Of Conversion
Type Of Error
polynomial
θ = 5040
Teff
polynomial
θ = 5040
Teff
= f(X − Y, [Fe/H])
= f(X − Y, [Fe/H])
standard deviation (in Kelvin)
for whole relation
standard deviation (in Kelvin)
for whole relation
polynomial
log Teff = f(X − Y )
error in percent
for whole relation
polynomial
Teff = f(X − Y )
polynomial
Teff = f(X − Y )
tabulated
polynomial
Teff = f(X − Y )
tabulated
standard deviation (in percent)
for whole relation
standard deviation of the
coefficients of the polynomial
∆ log Teff = 0.015 for F0 -- K51
∆Teff = 280 K for K7 -- M61
standard deviation of the
coefficients of the polynomial
∆ log Teff = 0.0151
polynomial
log Teff = f(X − Y, Fe/H, log g)
rms of residuals
4.2 Results
4 THESTATISTICALTESTWITHSINGLETTS
AAMR99g(8) (V-K)TCS
Fig. 1. Example for a plot of the differences D and DU as
well as the uncertainties U over spectral number SN for the
scale #8 of Alonso et al. (1999a) - AAMR99g(8) - for the
colour index (V − K)TCS. wTTS are marked with open cir-
cles, cTTS are marked with full circles. This scale was found
to fulfil our criteria, i. e. to be applicable to TTS.
4.1. Criteria for an applicable scale
In order to quantify discrepancies between TC and T S, the
uncertainty U of the differences D is calculated from the
quadratically added errors of both temperatures:
U = q∆2TCI + ∆2TS .
We account for the heterogeneous errors in the test and in
principle we regard the colour index temperature consistent
with the spectral type temperature if:
(2)
D < U
(3)
In order to perform statistical tests, a difference DU is now
defined which is the remaining difference when the uncer-
tainty U has already been subtracted from the absolute value
of D:
0
.
, if U > D
DU ≡ D − U , if U ≤ D and D > 0
DU ≡ D + U , if U ≤ D and D < 0
DU ≡
For each scale, the mean DU and the standard deviation
σ(DU ) is calculated. Furthermore, the differences D and
their uncertainties U are plotted versus spectral number SN
in order to calculate the slope of D(SN ) and to recognise pos-
sible systematic trends. An example of such a plot is given in
Fig. 1.
(4)
Finally, we used the following four criteria to decide
whether a scale is applicable to TTS and old stars, respec-
tively.
1. The mean difference DU should not be significantly dif-
ferent from zero. DU = 0 means that TS and TCI are
equal in average as is expected if both represent the effec-
tive temperature. A t-test with α = 0.01 and α = 0.05,
respectively, is done to check for significant differences.
2. According to Moore (1995), a t-test for less than 15 data
points is only valid if the data is normally distributed. As
this cannot be assumed a priori for our sample, only scales
yielding results for at least 15 stars are taken into account.
3. D(SN ) should have no significant slope, i. e. the slope
should be less than 3 σ different from constancy where σ
is the uncertainty estimate of the slope.
4. The mean error of the colour index scale ∆TCI should be
less than 500 K. This is rather a practical question than a
statistical criterion.
4.2. Results
4.2.1. Results for old stars
The detailed results are shown in Fig. 2. As expected, spectral
and colour index temperatures match quite well for the old
stars because both the conversion scales for spectral types and
for colour indices have been calibrated with data of old stars.
However, it is noteworthy that half of the scales do not pass
all four criteria although they are made for old stars.
Mean errors of the colour index scales larger than 500 K
appear only for the scales of Gratton et al. (1996) and
Houdashelt et al. (2000) who give errors of the coefficients of
their polynomial scales instead of errors of the whole scale.
At the same time, this legitimates the quite large error of
[Fe/H] (∆[Fe/H] = 0.55) adopted by us because the metal-
licity dependent scales of Alonso et al. (1996, 1999a) do not
show too large errors.
4.2.2. Results for TTS
We did the test for the TTS sample as a whole as well as sep-
arately for the cTTS and wTTS sub-samples. As an example,
the detailed results for one applicable scale are given in Fig.
1, the summarised results for all scales and the whole TTS
sample are given in Fig. 3. Due to the UV excess making
the stars bluer, the colour index temperatures for U − V and
B − V are in general larger than the mean spectral type tem-
peratures. For most of the U −V and B −V scales, the differ-
ences D = TCI −TS increase significantly with spectral num-
ber. This corresponds to a larger UV excess (in magnitudes)
for stars with weak photospheric UV emission (M stars) than
for stars with comparatively large photospheric UV emission
(earlier type stars). Analogously for V − L and V − M , the
colour index temperatures are too low due to the infrared ex-
cess.
For the colour indices V − H, V − K, IC − J, J − H,
and J − K at least some of the tested conversion scales pass
all four criteria, namely KH95d (V − H)JG, AAMR99g(8)
(V − K)TCS, GCC96g(1) (V − K)J, BLG94i (V − K)J,
"combined V − K", HSS94d IC − J, AAMR96d (J −
H)TCS, AAMR99g (J − H)TCS, AAMR96d (J − K)TCS,
and AAMR99g (J − K)TCS. However one has to keep in
mind that these criteria are only statistical ones -- individual
stars sometimes have differences DU of less than −1000 K
or more than +1000 K even if these "applicable scales" are
used.5 Nevertheless, as shown above the spectral type tem-
5 For the wTTS [HJS91] 4423, the spectral type used (M5) is
probably wrong because all colour index temperatures are more
than 1000 K larger than the mean spectral type temperature. The
observed colours are consistent on average with a spectral type of
about K2 to K3 (Kenyon and Hartmann 1995, table A5).
4 THESTATISTICALTESTWITHSINGLETTS
4.2 Results
scale index # of stars D
U and σ(D
) slope of D(SN) ∆T
CI
U
Fig. 2. The results of the statistical test for old stars. The first column gives the abbreviation of the used scale and the used
colour index. Scales which pass our test are marked bold. The second column gives the number of stars the scale yields a
result for. The third column gives a graphical representation of the DU (symbols) and of the σ(DU ) (error bars). Filled circles
mean that there is no significant difference. Open symbols denote a significant difference with a probability of error α = 0.01
(diamonds) and α = 0.05 (squares), respectively. The fourth column gives the slope of the D(SN ) if it is significant and the
fifth column the mean error ∆TCI of the colour index scale, if larger than 500 K. The unit of the slope is [K] since the abscissa
is the dimensionless spectral number (see Fig. 1). Further details are given in the text.
5.2 Thetestprocedure
5 THETESTWITHEVOLUTIONARYMODELS
perature of TTS is also not always a accurate measure for ef-
fective temperature. Thus our "applicable scales" can be used
only to verify the spectral temperature.
As wTTS have weaker variability and excesses than
cTTS, the mean differences DU as well as the σ(DU ) are in
general smaller for wTTS than for cTTS. Thus, if we consider
only wTTS, we find 25 applicable scales for B − V , V − J,
V − H, V − K, V − L, IC − J, J − H, and J − K. For
(V − R)C and (V − I)C only the combined scales pass our
criteria. If we consider only cTTS, we find only two applica-
ble scales (namely KH95d (V − H)JG and HSS94d IC − J).
The results for cTTS and wTTS are shown in Fig. 4 and Fig.
5, respectively.
by Steffen et al. (2001) themselves. For the eclipsing spectro-
scopic binary EK Cep, light curves for B, V , and R were ob-
tained by Ebbighausen (1966); Ebbighausen and Hill (1990);
and Khaliullin (1983)6.
Hill and Ebbighausen (1984) calculated colour indices
for the secondary from the relative contributions of both com-
ponents to the brightness in the B, V, and R band and the BVR
magnitudes of Khaliullin (1983) outside and within eclipse.
This object is not used as the colour indices calculated from
the primary eclipse are clearly discrepant from the ones cal-
culated from the values outside the eclipses.
5. The test with evolutionary models
Pre-main sequence evolutionary models give a clue to the un-
derstanding of the evolution of the very young stars. They
relate age and mass of a given star with other stellar parame-
ters, like for example the effective temperature. However, the
current understanding of the pre-main sequence evolution is
not sufficient in order to give predictions at the required level
of accuracy. In particular, there are free parameters in the un-
derlying physics which are not well-constrained, for example
the mixing-length parameter of the description of convection.
In this section we want to test combinations of evolution-
ary model and conversion scale for TTS where mass, TCI,
and either log g or Lbol is known.
5.1. Suitable test objects
In order to find suitable objects we first consider the most
fundamental parameter of a star, its mass. According to
Hillenbrand and White (2004), there are 18 TTS in 13 sys-
tems so far with known masses: six single cTTS for which
the mass could be derived from the mass of the surrounding
disc, three spectroscopic T Tauri binaries for which the incli-
nation i could be estimated, two eclipsing binaries with one
T Tauri component and two eclipsing T Tauri binaries.
The single cTTS cannot be used for our test as their ap-
parent bolometric luminosity can be determined only with
large uncertainties due to the excesses and the variability
of these objects. For example, GM Aur, one of the single
cTTS with known mass, has a luminosity of 1.00 L⊙ (ac-
cording to Valenti et al. 1993) or of 3.96 L⊙ (according to
Hillenbrand and White 2004).
The binaries can be used only for our test if resolved
colours of the components are given. But this is the case for
just two binaries, namely NTTS 045251+3016 and EK Cep.
Unfortunately we cannot use any of them as is explained in
the following.
For the astrometric T Tauri binary NTTS 045251+3016,
Steffen et al. (2001) give the resolved V − H colours which
were used to determine the temperature of the components
with the conversion scale of Kenyon and Hartmann (1995).
As we use only this scale to calculate colour index tempera-
tures with V − H, we would only reproduce the results given
Therefore, it seems that no suitable test object can be
found. On the other hand, the program "Nightfall" by R.
Wichmann allows calculating resolved colours from given
light curves. Thus, we could derive resolved colours for the
eclipsing wTTS binary V1642 Ori A because the BV JHK
light curves of this object were analysed with this program
by Covino et al. (2004) (see Table 2). We re-calculated their
final light curve solution using the adopted values T1 = 5200
K, q = 0.7305, and e = 0 as well as the values given in
Covino et al. (2004, table 4 [no-spots-solution] and table 5).
The resolved broad band magnitudes are given in Table 3.
Intrinsic colours were calculated using E(B − V ) = 0.10
as given by Covino et al. (2004). Our intrinsic colours are
consistent with the colours during secondary minimum ob-
tained by Covino et al. (2004).
Table 2. The physical parameters of the components of
V1642 Ori (Covino et al. 2004).
Primary V1642 Ori Aa
1.27 ± 0.01
1.44 ± 0.05
4.22 ± 0.02
5200 ± 150
M [M⊙]
R [R⊙]
logg [cgs]
Teff [K]
spectral type K1±1
Secondary V1642 Ori Ab
0.93 ± 0.01
1.35 ± 0.05
4.14 ± 0.02
4220 ± 150
K7-M0
Table 3. Resolved broad band magnitudes of V1642 Ori A
calculated by us using the software "Nightfall" by R. Wich-
mann. Only those values are given for which light curves ex-
ist. The errors are due to the errors of the "input data" (mass,
temperature, combined luminosity ..., see Table 2).
Primary V1642 Ori Aa
Secondary V1642 Ori Ab
B
V
J
H
K
13.74 ± 0.04
12.87 ± 0.04
11.19 ± 0.05
10.74 ± 0.06
10.60 ± 0.06
15.70 ± 0.16
14.44 ± 0.14
11.92 ± 0.10
11.17 ± 0.08
11.03 ± 0.08
5 THETESTWITHEVOLUTIONARYMODELS
5.3 Results
Although this refinement may be not really physical, it is
more precise than just interpolating mass and age by eye.
We assume that the interpolated isochrones have an error of
0.5 Myrs.
In the example shown in Fig. 6, we use the stellar param-
eters given in Table 2. One can see that the effective temper-
ature and luminosity of only the secondary is consistent with
the values predicted by the evolutionary model. The interpo-
lated isochrones give a coeval solution with an age of 9 to 16
Myrs.
The
following
evolutionary models were
(1998); D'Antona and Mazzitelli
used:
Baraffe et al.
(1994)
with "Alexander" opacities; D'Antona and Mazzitelli (1997)
with Y = 0.26 and XD = 2 · 10−5 or XD = 4 · 10−5, respec-
tively, as well as with Y = 0.28 and XD = 1 · 10−5,
XD = 2 · 10−5,or XD = 4 · 10−5, respectively;
Palla and Stahler
(2000); Yi et al.
(2003) with Z = 0.01, Z = 0.02, Z = 0.04, Z = 0.06, and
Z = 0.08.
(1999); Siess et al.
We used only those conversion scales which (a) were
found to be applicable for wTTS as described above and (b)
yield results for both components. This is the case for KH95d
(V − J)JG, KH95d (V − H)JG, HBS00d (V − K)CIT,
HBS00g (V − K)CIT, HBS00d (V − K)JG, HBS00g (V −
K)JG, Ben98i(FGK) (V − K)J, KH95d (V − K)JG, and
"combined V − K".
5.3. Results
5.3.1. Test with the HRD
The comparison in the HR diagram offers no really mean-
ingful test, mainly due to the large relative error of the lu-
minosities. All evolutionary models give consistent results in
combination with at least one conversion scale. As well, ev-
ery conversion scale gives consistent results in combination
with at least one evolutionary model. We get an overall range
of possible ages from 2 to 51 Myrs. In combination with ev-
ery conversion scale used within this test we obtain consistent
results for the models of Baraffe et al. (1998) with α = 1.5
and α = 1.9, of D'Antona and Mazzitelli (1994) with FST
mixing theory, all models of Siess et al. (2000) except for
Z = 0.10, and of Yi et al. (2003) with Z = 0.04.
The different evolutionary models yield different ages.
The models of D'Antona and Mazzitelli (1997) with Y =
0.26 give the lowest ages, namely 3 to 8 Myrs if the tem-
peratures given by Covino et al. (2004) are used. Using the
same temperatures, the models of Baraffe et al. (1998) with
α = 1.0 and Y = 0.275 yield the highest ages, namely 15 to
26 Myrs.
5.3.2. Test with the Kiel diagram
In the Kiel diagram the conversion scales KH95d (V −H)JG,
Ben98i(FGK) (V − K)J, and KH95d (V − K)JG as well
as the temperatures given by Covino et al. (2004) yield in-
consistent results for every evolutionary model. As well, the
evolutionary models Baraffe et al. (1998) with α = 1.0 and
α = 1.5, Palla and Stahler (1999), Siess et al. (2000), and
Fig. 6. The HR diagram with the components of V1642 Ori
A and tracks and isochrones of Palla and Stahler (1999).
The error boxes represent the parameters of V1642 Ori Aa
and V1642 Ori Ab, respectively. The effective temperatures
are adopted from Covino et al. (2004).
The solid lines are tracks interpolated for the limits on the
mass of these both objects. The interpolated isochrones give
the minimum and maximum age for both primary (dashed
line, 9 to 21 Myrs) and secondary component (dotted line, 3
to 16 Myrs).
5.2. The test procedure
We compare the masses, radii, and luminosities of the com-
ponents of V1642 Ori with the predictions of different evolu-
tionary models in both the HR diagram and the so-called Kiel
diagram (surface gravity vs. effective temperature). Thereby
the comparison is not only done with the effective tempera-
tures given by Covino et al. (2004). Instead we repeated the
comparison with several sets of temperatures resulting from
the application of colour index scales to the inferred colours
since neither the best evolutionary models nor the best colour
index scale is known. Of course this comparison does not al-
low one to draw conclusions on a certain scale or a certain
model but only on combinations of scales and models.
Each component is represented by an error box in the
HRD- and Kiel diagram, respectively. If (a) these rectangles
intersect with the tracks for the upper and lower limit of the
mass of the particular component and (b) an isochrone can
be found intersecting both rectangles, then the considered
combination of conversion scale and evolutionary model is
assumed to be consistent with the observational data given
by Covino et al. (2004). In order to better constrain possi-
ble masses and ages we interpolated the quite coarse grid
of tracks and isochrones given by the evolutionary models.
6 From the Julian dates given in table 1 of Khaliullin (1983)
(3902.3622 to 4629.5497), probably only 2 440 000 was subtracted
-- not 2 444 000 as given by Khaliullin (1983). Otherwise the light
curve would have been obtained from 1990 till 1991 -- too late for a
paper from 1983.
REFERENCES
REFERENCES
Yi et al. (2003) yield inconsistent results for every conversion
scale. For the remaining combinations of evolutionary mod-
els and conversion scales, only few yield consistent results
(see Table 4). The test with the Kiel diagram is more deciding
because the relative errors of the surface gravities are smaller
than the relative errors of the bolometric luminosities.
The overall range of possible ages (4 to 8 Myrs) is smaller
than the range obtained with the HRD.
It is remarkable that any evolutionary model does not
give consistent results if the temperatures by Covino et al.
(2004) are used in the Kiel diagram. All the suitable scales
except KH95d (B − V ) yield nearly the same primary tem-
perature as Covino et al. (2004). On the other hand, the
suitable scales imply higher secondary temperatures than
Covino et al. (2004) -- while all other scales do not.
6. Conclusions
We compiled several conversion scales which allow to derive
effective temperatures from broad band colour indices, in or-
der to examine their applicability to TTS.
These scales were first tested with a large sample of ap-
parently single TTS. For this purpose we used four statisti-
cal criteria. As a result, we found ten scales for V − H and
V − K as well as for the infrared colours IC − J, J − H, and
J − K which are consistent with the temperatures derived
from spectral type.
Furthermore we compared the colour index temperatures
and the dynamically derived masses of the components of
the eclipsing T Tauri binary V1642 Ori A with predictions of
pre-main sequence evolutionary models (Baraffe et al. 1998;
D'Antona and Mazzitelli 1994, 1997; Palla and Stahler 1999;
Siess et al. 2000; Yi et al. 2003), both in the HR diagram and
the Kiel diagram.
In the HR diagram all evolutionary models give consistent
results in combination with at least one conversion scale. As
well, every conversion scale gives consistent results in com-
bination with at least one evolutionary model. In the more de-
cisive Kiel diagram, the evolutionary models of Baraffe et al.
(1998) with α = 1.9, D'Antona and Mazzitelli (1994), and
D'Antona and Mazzitelli (1997) yield consistent results in
combination with at least some conversion scales. In this dia-
gram the scales HBS00d (V − K)CIT, HBS00d (V − K)JG,
and "combined V − K" appear to be most suitable. But it is
important to keep in mind that only combinations of evolu-
tionary model and conversion scale are tested -- neither evo-
lutionary models nor conversion scales alone.
As the Kiel diagram offers stricter constraints on evolu-
tionary models than the H-R diagram, we recommend to use
the Kiel diagram whenever possible.
Acknowledgements. We thank R. Neuhauser for giving us his com-
pilation of data of TTS, E. Covino for providing us the light curves
of V 1642 Ori A, and R. Wichmann for his program "Nightfall".
This research has made use of the SIMBAD database, operated
at CDS, Strasbourg, France and of NASA's Astrophysics Data Sys-
tem.
The comments of the anonymous referee helped to improve the
References
Alonso, A., Arribas, S., Mart´ınez-Roger, C.: 1994, A&AS 107, 365
Alonso, A., Arribas, S., Mart´ınez-Roger, C.: 1996, A&A 313, 873
Alonso, A., Arribas, S., Mart´ınez-Roger, C.: 1998, A&AS 131, 209
Alonso, A., Arribas, S., Mart´ınez-Roger, C.: 1999a, A&AS 140, 261
Alonso, A., Arribas, S., Mart´ınez-Roger, C.: 1999b, A&AS 139, 335
Ammler, M., Joergens, V., Neuhauser, R.: 2005, A&A 440, 1127
Baraffe, I., Chabrier, G., Allard, F., Hauschildt, P.H.: 1998, A&A
337, 403
Basri, G., Marcy, G.W.: 1994, ApJ 431, 844
Beichman, C.A., Boulanger, F., Moshir, M.: 1992, ApJ 386, 248
Bell, R., Gustafsson, B.: 1989, MNRAS 236, 653
Bessell, M.S.: 1979, PASP 91, 589
Bessell, M.S.: 1991, AJ 101, 662
Bessell, M.S., Brett, J.M.: 1988, PASP 100, 1134
Blackwell, D., Lynas-Gray, A.: 1994, A&A 282, 899
Briceno, C., Calvet, N., Gomez, M., Hartmann, L.W., Kenyon, S.J.,
Whitney, B.A.: 1993, PASP 105, 686
Briceno, C., Calvet, N., Kenyon, S., Hartmann, L.: 1999, AJ 118,
1354
Briceno, C., Hartmann, L., Stauffer, J., Mart´ın, E.: 1998, AJ 115,
2074
Cohen, M., Kuhi, L.V.: 1979, ApJ 227, L105+
Covino, E., Frasca, A., Alcal´a, J.M., Paladino, R., Sterzik, M.F.:
2004, A&A 427, 637
D'Antona, F., Mazzitelli,
I.: 1994, ApJS 90, 467, URL
http://www.mporzio.astro.it/{$\sim$}dantona/prems94.html
D'Antona,
Societa
http://www.mporzio.astro.it/{$\sim$}dantona/prems.html
1997, Memorie
807,
F., Mazzitelli,
Astronomica
della
URL
Italiana
68,
I.:
de Jager, C., Nieuwenhuijzen, H.: 1987, A&A 177, 217
di Benedetto, G.: 1998, A&A 339, 858
Ebbighausen, E.G.: 1966, AJ 71, 642
Ebbighausen, E.G., Hill, G.: 1990, A&AS 82, 489
Elias, J.H.: 1978, ApJ 224, 857
Fiorucci, M., Munari, U.: 2003, A&A 401, 781, URL
http://ulisse.pd.astro.it/Astro/ADPS/ReadMe/index.html
Fuhrmann, K.: 2004, Astronomische Nachrichten 325, 3
Gizis, J.E., Reid, I.N., Monet, D.G.: 1999, AJ 118, 997
Gomez, M., Jones, B.F., Hartmann, L., Kenyon, S.J., Stauffer, J.R.,
Hewett, R., Reid, I.N.: 1992, AJ 104, 762
Gratton, R., Carretta, E., Castelli, F.: 1996, A&A 314, 191
Haas, M., Leinert, C., Zinnecker, H.: 1990, A&A 230, L1
Hartigan, P.: 1993, AJ 105, 1511
Hartigan, P., Strom, K.M., Strom, S.E.: 1994, ApJ 427, 961
Hartmann, L., Stauffer, J.R., Kenyon, S.J., Jones, B.F.: 1991,
AJ 101, 1050
Hauck, B., Mermilliod, M.: 1998, A&AS 129, 431
Herbig, G.H., Bell, K.R.: 1988, Catalog of emission line stars of
the orion population : 3 : 1988 (Lick Observatory Bulletin, Santa
Cruz: Lick Observatory, -- c1988)
Herbig, G.H., Vrba, F.J., Rydgren, A.E.: 1986, AJ 91, 575
Herbst, W.,
Grossman,
Herbst,
instein,
ftp://www.astro.wesleyan.edu/pub/ttauri/
1994,
1906,
108,
D.:
D.K.,
AJ
E.J., We-
URL
Hill, G., Ebbighausen, E.G.: 1984, AJ 89, 1256
Hillenbrand, L.A., White, R.J.: 2004, ApJ 604, 741
Houdashelt, M., Bell, R., Sweigart, A.: 2000, AJ 119, 1448
Johnson, H.L.: 1966, ARA&A 4, 193
Jones, B.F., Herbig, G.H.: 1979, AJ 84, 1872
Kenyon, S.J., Calvet, N., Hartmann, L.: 1993, ApJ 414, 676
Kenyon, S.J., Gomez, M., Marzke, R.O., Hartmann, L.: 1994a,
manuscript substantially.
AJ 108, 251
REFERENCES
REFERENCES
Table 4. Results of the test with the Kiel diagram for V1642 Ori A. The combinations of evolutionary models (rows) and
conversion scales (columns) yielding consistent results are marked with "X".
"combined V − K" refers to the mean temperature of all V − K scales (see section 2.3). The parameters of the evolutionary
models are as follows: α -- mixing length parameter, FST -- full spectrum of turbulence, MLT -- mixing length theory, Y --
relative Helium abundance, XD -- relative Deuterium abundance.
KH95d
HBS00d
HBS00g
HBS00d
HBS00g
"combined V − K"
(V − J)JG
(V − K)CIT
(V − K)CIT
(V − K)JG
(V − K)JG
X
Baraffe et al. (1998)
α = 1.9
D'Antona and Mazzitelli (1994)
FST, "Alexander" opacities
D'Antona and Mazzitelli (1994)
MLT, "Alexander" opacities
D'Antona and Mazzitelli (1997)
Y = 0.26, XD = 2 · 10−5
D'Antona and Mazzitelli (1997)
Y = 0.26, XD = 4 · 10−5
D'Antona and Mazzitelli (1997)
Y = 0.28, XD = 1 · 10−5
D'Antona and Mazzitelli (1997)
Y = 0.28, XD = 2 · 10−5
D'Antona and Mazzitelli (1997)
Y = 0.28, XD = 4 · 10−5
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
Kenyon, S.J., Hartmann, L.: 1995, ApJS 101, 117
Kenyon, S.J., Hartmann, L., Hewett, R., et al.: 1994b, AJ 107, 2153
Kenyon, S.J., Hartmann, L.W., Strom, K.M., Strom, S.E.: 1990,
Skrutskie, M.F., Dutkevitch, D., Strom, S.E., Edwards, S., Strom,
K.M., Shure, M.A.: 1990, AJ 99, 1187
Steffen, A.T., Mathieu, R.D., Lattanzi, M.G., et al.: 2001, AJ 122,
AJ 99, 869
997
Khaliullin, K.F.: 1983, AZh 60, 72
Leinert, C., Haas, M.: 1989, ApJ 342, L39
Leinert, C., Haas, M., Mundt, R., Richichi, A., Zinnecker, H.: 1991,
A&A 250, 407
Leinert, C., Zinnecker, H., Weitzel, N., Christou, J., Ridgway, S.T.,
Strom, K.M., Strom, S.E.: 1994, ApJ 424, 237
Strom, K.M., Strom, S.E., Edwards, S., Cabrit, S., Skrutskie, M.F.:
1989, AJ 97, 1451
Strom, K.M., Strom, S.E., Wilkin, F.P., et al.: 1990, ApJ 362, 168
Torres, C.A.O., Quast, G., de La Reza, R., Gregorio-Hetem, J., Lep-
Jameson, R., Haas, M., Lenzen, R.: 1993, A&A 278, 129
ine, J.R.D.: 1995, AJ 109, 2146
Luhman, K.L.: 1999, ApJ 525, 466
Magazzu, A., Martin, E.L.: 1994, A&A 287, 571
Martin, E.L., Rebolo, R., Magazzu, A., Pavlenko, Y.V.: 1994, A&A
Valenti, J.A., Basri, G., Johns, C.M.: 1993, AJ 106, 2024
Vrba, F.J., Rydgren, A.E., Zak, D.S.: 1985, AJ 90, 2074
Walter, F.M., Brown, A., Mathieu, R.D., Myers, P.C., Vrba, F.J.:
282, 503
Moneti, A., Zinnecker, H.: 1991, A&A 242, 428
Moore, D.S.: 1995, Basic practice of statistics (Freeman & Com-
pany)
1988, AJ 96, 297
Yi, S.K., Kim, Y., Demarque, P.: 2003, ApJS 144, 259
Moro, D., Munari, U.:
2000, A&AS 147,
361, URL
http://ulisse.pd.astro.it/Astro/ADPS/ReadMe/index.html
Myers, P.C., Fuller, G.A., Mathieu, R.D., Beichman, C.A., Benson,
P.J., Schild, R.E., Emerson, J.P.: 1987, ApJ 319, 340
Ochsenbein, F., Halbwachs, J.L.: 1987, Bulletin d'Information du
Centre de Donnees Stellaires 32, 83
Palla, F., Stahler, S.W.: 1999, ApJ 525, 772
Perrin, G., du Foresto, V.C., Ridgway, S.T., Mariotti, J., Traub, W.,
Carleton, N., Lacasse, M.: 1998, A&A 331, 619
Reid, I.N., Hawley, S.L.: 1999, AJ 117, 343
Rydgren, A.E.: 1984, Publications of the U.S. Naval Observatory
Second Series 25, 1
Samus, N.N., Durlevich, O.V., et al.: 2004, VizieR Online Data Cat-
alog 2250, 0
Schmidt-Kaler: 1982, in K.H. Hellwege, ed., Landolt-Boernstein,
Group VI, 449 -- 456 (Springer Verlag Berlin)
Sekiguchi, M., Fukugita, M.: 2000, AJ 120, 1072
Selby, M.J., Hepburn, I., Blackwell, D.E., Booth, A.J., Haddock,
D.J., Arribas, S., Leggett, S.K., Mountain, C.M.: 1988, A&AS
74, 127
Siess, L., Dufour, E., Forestini, M.: 2000, A&A 358, 593, URL
http://www-astro.ulb.ac.be/{$\sim$}siess/database.html
Simon, M., Chen, W.P., Howell, R.R., Benson, J.A., Slowik, D.:
1992, ApJ 384, 212
Simon, M., Ghez, A.M., Leinert, C., et al.: 1995, ApJ 443, 625
REFERENCES
REFERENCES
scale index # of stars D
U and σ(D
) slope of D(SN) ∆T
CI
U
Fig. 3. Same as Fig. 2, but for TTS.
REFERENCES
REFERENCES
scale index # of stars D
U and σ(D
) slope of D(SN) ∆T
CI
U
Fig. 4. Same as Fig. 2, but for cTTS.
REFERENCES
REFERENCES
scale index # of stars D
U and σ(D
) slope of D(SN) ∆T
CI
U
Fig. 5. Same as Fig. 2, but for wTTS.
|
astro-ph/0312615 | 1 | 0312 | 2003-12-24T15:43:35 | High-Energy Emission from Presupernovae | [
"astro-ph"
] | The non-thermal emission in the magnetospheres of presupernova collapsing stars with initial dipole magnetic fields and a certain initial energy distribution of the charged particles in a magnetosphere is considered. The analysis of particle dynamics and its emission in the stellar magnetosphere under collapse show that the collapsing stars can be the powerful sources of a non-thermal radiation produced by the interaction of charged particles with the magnetic field. The radiation flux grows with decreasing stellar radius. This flux can be observed in the form of radiation burst with duration equal to the stellar collapse time. The radiation flux depends on the distance to the stars, the particle spectrum and magnetic field in the magnetosphere. The radiation flux is calculated for various collapsing stars with initial dipole magnetic fields and an initial power series, relativistic Maxwell, and Boltzmann particle energy distribution in the magnetosphere. This radiation can be observed by means of modern astronomical instruments. | astro-ph | astro-ph | HIGH-ENERGY EMISSION FROM PRESUPERNOVAE
Volodymyr Kryvdyk1
1 Depart. of Astronomy, Kyiv University, av. Glushkova 6, Kyiv 03022, Ukraine
ABSTRACT
The non-thermal emission in the magnetospheres of presupernova collapsing stars with initial dipole magnetic
fields and a certain initial energy distribution of the charged particles in a magnetosphere is considered. The
analysis of particle dynamics and its emission in the stellar magnetosphere under collapse show that the
collapsing stars can be the powerful sources of a non-thermal radiation produced by the interaction of charged
particles with the magnetic field. The radiation flux grows with decreasing stellar radius. This flux can be
observed in the form of radiation burst with duration equal to the stellar collapse time. The radiation flux
depends on the distance to the stars, the particle spectrum and magnetic field in the magnetosphere. The
radiation flux is calculated for various collapsing stars with initial dipole magnetic fields and an initial power
series, relativistic Maxwell, and Boltzmann particle energy distribution in the magnetosphere. This radiation
can be observed by means of modern astronomical instruments.
INTRODUCTION
There are three ways to observe the stars on the stage of gravitational collapse, namely to detect 1) neutrinos,
2) the gravitational waves, and 3) electromagnetic radiation, arisen by collapse.
The first neutrino signal from supernova stars was detected in 1987 from SN 1987 A (Bionta, et al., 1987,
Hirata , et al., 1987). It is the unique observation of a star on the stage of its explosion.
The stars must emit electromagnetic and gravitational waves under the collapse. But these electromagnetic
waves are very low frequencies (Cunningam , et al. , 1978, 1979, 1980, Henriksen, et al.,1979, Moncrief ,1980)
therefore their can not observed near Earth. Still, these waves are not detected, and maybe this is the main
problem in the theory of stellar collapse.
In this paper the one with method for the observation of the stellar collapse is proposed using the radiation
arisen in magnetospheres of collapsing stars. This radiation will be generated when the star magnetosphere
compress under the collapse and its magnetic field considerably increases. Thus a cyclic electric field is
produced, involving of acceleration charged particles that generate radiation when moving in the magnetic field.
The frequencies of this radiation are very high (from gamma-rays to radio waves) therefore they can be
observed near Earth
COLLAPSAR MAGNETOSPHERE
We considered the collapse of stars with an initial dipolar magnetic field and three typical particles
EK
EN
)
(1
γ−
=
the magnetosphere: 1) power-series
distributions
in
; 2)
relativistic Maxwell
p
kTE
EK
EN
kTE
EN
K
)
(
exp(
)
/
/
)
(3
exp(
)
2
=
−
=
−
; and 3) Boltzmann distributions
B
M
2
The external electromagnetic field of the collapsing stars will change as (Ginzburg and Ozernoy, 1964)
Br
r
t
2 3
)(
cos
−
θ
=
µ
,
B
t
r
sin)(
3
−=
µ
θ
,
θ
0=ϕB
(1)
µ
∂
E
rc
1
−
t
∂
−=
sin
−
2
ϕ
θ
,
E r
= θE
0=
,
where µ(t) is the variable magnetic momentum of the collapsar.
The field structure and particles dynamics in the magnetosphere are influenced by three factors: particles
pressure, collisions, and star rotation. As follow from a detailed analysis (Kryvdyk, 1999) these effects can be
neglected during the collapse.
In order to investigate the particle dynamics during the collapse we use the method of adiabatic invariant. In
this case the particles energy will change as results of the two mechanisms: a betatron acceleration in the
variable magnetic field, 2) bremsstrahlung energy losses in this field.
For the betatron acceleration in drift approximation we can write (Bakhareva and Tverskoj, 1981)
(
dE
dt
where p and v is particle impulse and particle velocity,
For the dipolar magnetic field (1)
(cid:71)
pvdivu
(1 / 3)
= −
)
a
(2)
(cid:71)(cid:71)
[2 BE
−
is the drift velocity.
cB
=
]
u
(
dE
dt
)
a
=
)3/4(
µ
1
−
(
∂µ
/
∂
t
)
pvf
)(
θ
4
θf
)(
31)(1
3(
cos
cos
)
θ
θ
=
−
+
where
.
The magnetic moment µ(t) of collapsar change in a results of the change their radius R(t) under the influence
of gravitational field according to the law of free fall
2
−
2
dR
dt
=
GM R
(
(2
*
−
1) /
RR
*
)
1 / 2
.
Here R* =R0/R.
Passing to the new variable R= R (t) we can write for the betatron acceleration
(
dE
dt
)
a
=
4
3
k
1
GM
2(
/
R
)
2/13
[(
R
−
RR
/)
0
0
]
2/1
f
)(
θ
E
Here k1 =2 and k1 =1 for relativistic and non-relativistic particles respectively.
For the bremsstrahlung energy losses (Ginzburg and Syrovatskij, 1964)
(
dE
dt
)
s
=
4
(
e
/ 6
m c B R g
)
)(
4 7
2
0
0
( ,
θ α
)
R E r
2
2
6
−
,
where α is the angle between the magnetic field and the particle velocity.
g θ α
Here
(1 3 cos
( ,
.
) sin
)
2
2
α
θ
= +
The resulting rate of particle energy change in the magnetosphere is
(
dE
dt
)
=
(
dE
dt
)
a
+
(
dE
dt
)
s
=
4
3
k GM R
(
(2
1
*
−
1) /
R R
*
)
3 1 / 2
f
( )
θ
E
−
4
(
e
/ 6
m c
4 7
)(
Φ
0
2
)
g
( ,
θ α
)
R E r
2
2
−
6
(3)
Particle dynamics can be investigated using the equation of transitions particle in the regular magnetic fields
(Ginzburg and Syrovatskij, 1964)
N
∂
t
∂
+
∂
E
∂
(
dEN
dt
) 0
=
For the new variable R= R(t) this equation becomes
Here
∂
∂
N
R
=
f E R
,
(
1
)
∂
∂
N
E
+
f E R N
)
,
(
2
=
0,
(4)
1
−
REf
A
ER
{
)
(
,
−
=
1
1
R
REf
A
2
(
,
)
{
1
−
=
−
2
1
A
fk
;3/)(
4
θ
=
1
1
A
cm
e
6/
(
4
4
7
=
2
)(
RB
0
RRRA
/(
[
3
0
2
RRR
/(
[
3
−
0
R
)]
2/1
−
R
)]
2/1
E
};
E
},
()
22
0
R
0
GM
2/
)
2/1
g
,(
αθ
)
Eq. (4) can not be solved in the general case and so two special cases are considered: (i) when energy losses do
not influence the particle spectrum spectrum in the magnetosphere and (ii) when the energy losses determine
the particle spectrum .
The solution of Eq. (4) in these two cases is given by
REN
,
)
(
i
1
REN
i
,
(
)
2
REN
i
)
,
(
3
=
=
=
β
1
REK
−
−
γ
;
p
*
kTE
REK
/
exp(
2
− β
−
2
M
*
RK
kTE
exp(
/
− β
−
;
3
B
*
)
(5)
REN
)
(
,
ii
1
N
RE
,
(
)
ii
=
RE
N
)
,
(
=
ii
3
2
K
1(
exp(
=
γ −
γ
−
p
EK
1(
exp(
2
γ−−
M
K
exp(
1(
)
γ−−
))
;
1
kTE
)
/
)
; (6)
1
kTE
)
/
.
B
1
A F R R r E
A E kT E
A E kT E
A
3);
(
1)
ln
/
(
ln
/
(
1);
;
)
,
(
6
−
−
=
−
=
−
=
=
β
β
γ
β
γ
Here
B
M
P
1
1
1
0
*
2
1
Eqs. (5) determine the particle spectrum in the magnetosphere and its evolution during the initial stage of the
collapse when the energy losses can be neglected. We will consider this case in this paper. Eqs. (6) determine
the particles spectrum on the final stage of the collapse, when the magnetic field attains an extreme value and
the energy losses influence the particle spectrum considerably. This case we will consider in a later paper.
NON-THERMAL EMISSION FROM COLLAPSAR
The ratio between the radiation flux from collapsar with radius R and its initial radiation flux (by R= R0 ) for
the power-series has been obtained previously in the paper ( Kryvdyk,1999)
I
2/
∞
∫ ∫
0
0
Similarly for relativistic Maxwell and Boltzmann distributions we can write down
dEd
θθ
/
(
νν
R
A
1
*
sin
R
γ
*
1(
γ
−
=
2/)
P
ν
P
ν
0
−
2
−
)2
/
I
)
0
(
γ
π
,
(7)
I
M
ν
/
I
M
ν
0
=
(
/
νν
3
)
R
−
*
0
/1(
kT
π
)
I
B
ν
/
I
B
ν
0
=
/
(
νν
3
)
R
−
*
(
kT
0
π
)
2/
∞
∫ ∫
0
0
2/
∞
∫ ∫
0
0
β
M
R
−
*
exp(
kTE
/
sin)
dEd
θθ
,
(8)
β
B
R
−
*
E
−
2
exp(
kTE
/
sin)
dEd
θθ
,
(9)
Using Eq. (7)-(9) the radiation flux from the collapsar magnetospheres with variable dipole magnetic fields can
be calculated. The ratio between the radiation flux from a collapsars with radius R and their initial flux by ν/ν0
=1 are:
1 ≤ IνP / IνP0 ≤ 1.34. 1010 for 2.4≤ γ ≤3.4, 10 ≤ R* ≤ 1000,
1 ≤ IνB / IνB0 ≤ 4.86.105 for 1 eV ≤ kT ≤ 9eV, 145 ≤ R* ≤ 850,
1 ≤ IνM / IνM0 ≤ 2.23.1011 for 1 eV ≤ kT ≤ 9eV, 145 ≤ R* ≤ 850.
These values obtained by the numerical integration of the equations (7)-(9) for 2 eV ≤ E ≤ 109 eV and the
different R* , kT, γ.
CONCLUSIONS
The magnetic fields of presupernova stars increase during the collapse. These variable magnetic fields
accelerate the charged particles in the magnetospheres of collapsing stars. These particles emit the
electromagnetic waves in the wide frequency band, from gamma rays to radio waves. The radiation flux grows
during collapse and to reaches a maximum on the final stage of collapse. This radiation can be observed as
bursts in the all-frequency band, from radio to gamma ray. The burst duration completes with the collapse time.
The radiation flux from collapsar on the final stage of collapse exceeds the initial flux a million times. Thus the
collapsar can be the powerful sources of the non-thermal radiation when before a supernova flares, the star
compresses and emits bursts. Where can we seek these bursts? First of all in powerful gamma bursts and X
bursts which are not periodical..
What problems can arise in the program for the astrophysical observation of bursts from presupernova
collapsing stars? Above all now the theories of the stellar evolution do not enable us to locate the presupernova
collapsar with enough accuracy. We can indicate only on the stars types that can collapse to supernova, but we
can not point to its location. This fact is principal problem in the observational astrophysics program for the
search of collapsing presupernova stars. A next problem is how to choose the impulse from presupernova
collapsar from the great number of the bursts with unknown origin.
REFERENCES
Bakhareva M.F., and B.A. Tverskoj, The particles energy variation in the variable interplanetary magnetic field,
Geomagn .i Aeron., 21, 401, 1981 (in Russian)
Bionta R.M., G. Blewitt, C.B Bratton., et al .Observation of a neutrino burst in coincidence
with supernova 1987A in Large Magellanic Cloud, Phys. Rev.Lett.,.-58, 1494-1496, 1987
Ginzburg V.L., and L.M. Ozernoy, About gravitational collapse of magnetic star,
Zh.. Exper. i Theor. Fiz.( ZhEThF) 47, 1030-1040, 1964 (In Russian)
Ginzburg V.L., and S.I. Syrovatskii, Origin of cosmic rays, Izd. AN USSR, Moscow,
1963 (In Russian)
Cunningam C.T., R.H. Price, and V. Moncrief , Radiation from collapsing relativistic
stars.I. Linearized, Astrophys. J., 224, 643-667, 1978.
Cunningam C.T., R.H. Price, and V. Moncrief, Radiation from collapsing relativistic
stars.II. Linearized even-parity radiation, Astrophys. J., 230, 870-892, 1979.
Cunningam C.T., R.H. Price, and V. Moncrief, Radiation from collapsing relativistic
stars. III. Second order perturbation of collapse with rotation, Astrophys. J., 236, 674-692,
1980.
Henriksen R.N., W.Y. Chau , and K.L.Chau , Magnetic dipole radiation from a
exploding or collapsing magnetised rotating spheroid , Astrophys. J, 227, 1013-1018, 1979.
Hirata K., T. Kajiata, M. Koshiba, et al. Observation of a neutrino burst from
the supernova 1987A , Phys. Rev.Lett., 58, 1490-1493, 1987
Kryvdyk V., Electromagnetic radiation from collapsing stars.I. Power- series
distribution of particles in magnetospheres, Mon. Not. R.Astron. Soc., 309, 593-598, 1999.
Moncrief V., Radiation from collapsing relativistic stars. IY. Black hole recoild, Astrophys.J,
238, 333-337, 1980.
|
0705.1622 | 1 | 0705 | 2007-05-11T11:06:14 | Triaxial Analytical Potential-Density Pairs for Galaxies | [
"astro-ph"
] | We present two triaxial analytical potential-density pairs that can be viewed as generalized versions of the axisymmetric Miyamoto and Nagai and Satoh galactic models. These potential-density pairs may be useful models for galaxies with box-shaped bulges. The resulting mass density distributions are everywhere non-negative and free from singularities. Also, a few numerically calculated orbits for the Miyamoto and Nagai-like triaxial potential are presented. | astro-ph | astro-ph |
Triaxial Analytical Potential-Density Pairs for
Galaxies
Daniel Vogt∗and Patricio S. Letelier†
Departamento de Matem´atica Aplicada-IMECC, Universidade Estadual
de Campinas, 13083-970 Campinas, S. P., Brazil
October 25, 2018
Abstract
We present two triaxial analytical potential-density pairs that can
be viewed as generalized versions of the axisymmetric Miyamoto and
Nagai and Satoh galactic models. These potential-density pairs may be
useful models for galaxies with box-shaped bulges. The resulting mass
density distributions are everywhere non-negative and free from sin-
gularities. Also, a few numerically calculated orbits for the Miyamoto
and Nagai-like triaxial potential are presented.
1
Introduction
There are several three-dimensional analytical models in the literature for
the gravitational field of different types of galaxies and galactic components.
Jaffe [1] and Hernquist [2] discuss models for spherical galaxies and bulges.
Three-dimensional models for flat galaxies were obtained by Miyamoto and
Nagai [3] and Satoh [4]; de Zeeuw and Pfenniger [5] presented a sequence
of infinite triaxial potential-density pairs relevant for galaxies with massive
haloes of different shapes. Long and Murali [6] derived simple potential-
density pairs for a prolate and a triaxial bar by softening a thin needle with
a spherical potential and a Miyamoto and Nagai potential, respectively. See
[7] for a discussion on other galactic models. There also exist several general
relativistic models of disks, e. g., [8] -- [15]. A general relativistic version of
the Miyamoto and Nagai models was studied by Vogt and Letelier [16].
∗e-mail: [email protected]
†e-mail: [email protected]
1
Although axial symmetry is a common approximation to many disk
galactic models, recent statistics of bulges of disk galaxies (S0 -- Sd) have
revealed that nearly half of them are box- and peanut- shaped [17]. About
4% of all galaxies with box- and peanut- shaped bulges have so called "Thick
Boxy Bulges": they are box shaped and are large with respect to the di-
ameters of their galaxies [18]. In this work we consider two simple triax-
ial potential-density pairs whose mass density distributions are box-shaped
along one axis. The first is constructed by applying on each Cartesian coor-
dinate a transformation similar to that used by Miyamoto and Nagai; this
is done in subsection 2.1. The second pair is a triaxial generalization of one
studied by Satoh, and is presented in subsection 2.2. In section 3 some orbits
for the Miyamoto and Nagai-like triaxial potential are exhibited. Finally,
the results are discussed in section 4.
2 Triaxial Models for Galaxies
In the following potential-density pairs, the mass-density distribution is ob-
tained directly from Poisson equation
ρ =
1
4πG
(Φ,xx + Φ,yy + Φ,zz) ,
(1)
where Φ(x, y, z) is the gravitational potential.
2.1 Triaxial Miyamoto and Nagai-like Model 1
We start with the gravitational monopole potential in Cartesian coordinates,
Φ = −
Gm
,
px2 + y2 + z2
and apply the transformations x → a1 +px2 + b2
z → a3 + pz2 + b2
equation (1), we obtain the following density distribution
2 and
3, where the ai, bi are non-negative constants. Using
1, y → a2 +py2 + b2
(2)
1
¯ρ =
4πξ3η3χ3 [(¯a1 + ξ)2 + (¯a2 + η)2 + (1 + χ)2]5/2 (cid:8)¯b2
+¯a1(¯a2 + η)2 + ¯a1(1 + χ)2(cid:3) + ¯b2
+¯a2(1 + χ)2(cid:3) + ¯b2
3ξ3η3(cid:2)(1 + χ)2(1 + 3χ) + (¯a1 + ξ)2 + (¯a2 + η)2(cid:3)(cid:9) ,
2ξ3χ3(cid:2)(¯a2 + η)2(¯a2 + 3η) + ¯a2(¯a1 + ξ)2
1η3χ3(cid:2)(¯a1 + ξ)2(¯a1 + 3ξ)
(3)
2
(a)
0.01
0.002
0.001
2.5
5
0.005
0
x−
(c)
(b)
5
2.5
z−
0
−2.5
−5
−5
−2.5
0
y−
2.5
5
5
2.5
y−
0
−2.5
−5
−5
−2.5
5
2.5
z−
0
−2.5
−5
−5
−2.5
0
x−
2.5
5
Figure 1: Constant density curves of equation (3) on the planes (a) ¯z = 0,
(b) ¯x = 0 and (c) ¯y = 0 with parameters ¯a1 = 1.0, ¯b1 = 1.0, ¯a2 = 0.5,
¯b2 = 1.0 and ¯b3 = 0.5. The contour levels in (b) and (c) are the same as
shown in (a).
2 and χ = q¯z2 + ¯b2
3.
1,
where the variables and parameters are rescaled in terms of a3: ¯x = x/a3,
¯y = y/a3, ¯z = z/a3, ¯ai = ai/a3, ¯bi = bi/a3, ¯ρ = a3
η = q¯y2 + ¯b2
3ρ/m, and ξ = q¯x2 + ¯b2
The density distribution equation (3) is always non-negative and free
from singularities. For ¯a1 = ¯a2 = 0 the potential-density pair is axisymmet-
ric with respect to the z axis, and in particular if ¯b1 = ¯b2 = 0 we recover the
Miyamoto & Nagai model 1. Figures 1(a) -- (c) show some isodensity curves
of equation (3) on the planes (a) ¯z = 0, (b) ¯x = 0 and (c) ¯y = 0 with
parameters ¯a1 = 1.0, ¯b1 = 1.0, ¯a2 = 0.5, ¯b2 = 1.0 and ¯b3 = 0.5. From a
top view, the matter distribution is box-like shaped, whereas from a lateral
view matter seems more flattened in a disk-like manner. Plotting some iso-
density curves with other values of the parameters allows us conclude that
3
larger values of ¯a1 and ¯a2 lead to larger deviations from axisymmetry, and
the degree of flatness with respect to the plane ¯z = 0 depends on ¯b3 as in
the original Miyamoto and Nagai models.
2.2 A Triaxial Satoh-like Model
Satoh [4] derived a family of three-dimensional axisymmetric mass distri-
butions by flattening the higher order Plummer models of order n. When
n → ∞ the potential takes a rather simple form
Gm
Φ = −
hx2 + y2 + z2 + a(a + 2√z2 + b2)i1/2 .
(4)
We propose the following triaxial generalization of the above potential:
Φ = −Gm(cid:20)x2 + a1(a1 + 2qx2 + b2
1) + y2 + a2(a2 + 2qy2 + b2
+a3(a3 + 2qz2 + b2
3)(cid:21)−1/2
2) + z2
.
(5)
The corresponding mass density distribution follows from equation (1)
1
¯ρ =
4πξ3η3χ3 [(¯x2 + ¯a1(¯a1 + 2ξ) + ¯y2 + ¯a2(¯a2 + 2η) + ¯z2 + 1 + 2χ]5/2 ×
(cid:8)¯a1¯b2
1η3χ3(cid:2)¯x2 + ¯y2 + ¯z2 + (¯a1 + 2ξ)(¯a1 + 3ξ) + ¯a2(¯a2 + 2η) + 1 + 2χ(cid:3)
+¯a2¯b2
2ξ3χ3(cid:2)¯x2 + ¯y2 + ¯z2 + (¯a2 + 2η)(¯a2 + 3η) + ¯a1(¯a1 + 2ξ) + 1 + 2χ(cid:3)
+¯b2
3ξ3η3(cid:2)¯x2 + ¯y2 + ¯z2 + (1 + 2χ)(1 + 3χ) + ¯a1(¯a1 + 2ξ) + ¯a2(¯a2 + 2η)(cid:3)(cid:9) ,
(6)
where the variables and parameters were rescaled as in subsection 2.1. The
density distribution equation (6) is also non-negative and free from singu-
larities. For ¯a1 = ¯a2 = 0 we recover the original Satoh model. In figures
2(a) -- (c) we display some isodensity curves of equation (6) on the planes (a)
¯z = 0, (b) ¯x = 0 and (c) ¯y = 0 with parameters ¯a1 = 1.0, ¯b1 = 1.0, ¯a2 = 0.5,
¯b2 = 1.0 and ¯b3 = 0.5.
3 Orbits in triaxial potential-density pairs
In this section we briefly discuss some orbits calculated numerically for the
triaxial Miyamoto and Nagai-like potential-density pair presented in subsec-
tion 2.1. We first consider orbits on the z = 0 plane. Since the potential
4
5
2.5
y−
0
−2.5
(a)
0.1
0.2
0.05
0.02
(b)
5
2.5
z−
0
−2.5
−5
−5
−2.5
0
x−
(c)
2.5
5
−5
−5
−2.5
0
y−
2.5
5
5
2.5
z−
0
−2.5
−5
−5
−2.5
0
x−
2.5
5
Figure 2: Constant density curves of equation (6) on the planes (a) ¯z = 0,
(b) ¯x = 0 and (c) ¯y = 0 with parameters ¯a1 = 1.0, ¯b1 = 1.0, ¯a2 = 0.5,
¯b2 = 1.0 and ¯b3 = 0.5. The contour levels in (b) and (c) are the same as
shown in (a).
5
(a)
(b)
4.5
4.0
−
R
3.5
3.0
2.5
0
1
2
3
ϕ/(2π)
4
5
6
8
6
4
−
R
2
0
1
2
ϕ/(2π)
3
Figure 3: Radial coordinate ¯R = R/a3 as function of polar coordinate ϕ for
motion with triaxial potential (solid curves) and with axisymmetric potential
(dashed curves). Parameters: ¯a1 = 1.0, ¯b1 = 1.0, ¯a2 = 0.5, ¯b2 = 1.0 and
¯R = 0, ϕ = 0 and ϕ was calculated for
¯b3 = 0.5. Initial conditions: ¯R = 3.0,
an energy ¯E = −0.15 in (a) and ¯E = −0.10 in (b).
is a function of the polar angle, there is no conservation of the z compo-
nent of the angular momentum. In figures 3(a) -- (b) we compare the orbits
for the triaxial potential (solid curves) and for the axisymmetric potential
(dashed curves). The parameters used were the same as in figure 1 and for
the axisymmetric case ¯a1 = ¯a2 = 0. In the equations of motion time was
rescaled as ¯t = (Gm/a3
3)1/2t and the conserved mechanical energy per unit
mass E was rescaled as ¯E = (Gm/a3)−1E. Initial conditions in cylindrical
¯R = 0, ϕ = 0 and ϕ was determined for a
coordinates were set as ¯R = 3.0,
total energy ¯E = −0.15 in figure 3(a) and ¯E = −0.10 in figure 3(b). In the
axisymmetric case the orbits shown are precessing ellipses. When axisym-
metry is destroyed, so is the regular oscillation of the radial coordinate. At
a higher energy, both orbits are qualitatively almost identical.
Figures 4(a) -- (d) show an example of three-dimensional orbit for (a) ax-
isymmetric potential and (b) triaxial potential. In figures 4(c) -- (d) the radial
coordinate ¯R and the coordinate ¯z, respectively, are plotted as function of
time ¯t for the triaxial case (solid curves) and axisymmetric case (dashed
curves). Here the parameters used were the same as in the previous exam-
¯R = 0, ϕ = 0 and the imposed
ple; the initial conditions were ¯R = 3.0,
initial constrait ¯z = 0.5 ¯R ϕ was calculated for an energy of ¯E = −0.15. As
before, motion with triaxial potential is more irregular compared with the
axisymmetric one. This reflects the loss of one of the motion integrals, the
angular momentum.
6
(a)
(b)
1.5
0.75
z−
0
−0.75
−1.5
4
2
y−
0
−2
−4
−4
−2
0
x−
4
2
2
y−
0
−2
−4
−4
−2
0
x−
4
2
(c)
(d)
1.5
0.75
z−
0
−0.75
−1.5
0
50
100
t−
150
200
250
50
100
t−
150
200
250
1.5
0.75
z−
0
−0.75
−1.5
4
4
3.5
−
R
3
2.5
0
Figure 4: Three-dimensional orbit for (a) axisymmetric potential and (b)
triaxial potential. (c) Radial coordinate ¯R as function of time ¯t. (d) Co-
ordinate ¯z as function of time ¯t. Solid curves represent motion for triaxial
potential and dashed curves the axisymmetric case. The parameters are the
same as in figure 3. Initial conditions: ¯R = 3.0,
¯z = 0.5 ¯R ϕ
and energy ¯E = −0.15.
¯R = 0, ϕ = 0,
7
4 Discussion
By applying a Miyamoto and Nagai transformation on the three Carte-
sian coordinates of the monopole potential, we obtained a triaxial potential-
density pair whose mass density distribution is everywhere non-negative
and free from singularities; also, it is box-shaped with respect to the z
axis. A triaxial version of one of Satoh's axisymmetric models also yields a
mass-density distribution with similar characteristics. We believe that these
simple analytical models may be useful for disk galaxies having box-shaped
bulges. Some numerically calculated orbits for the Miyamoto and Nagai-like
triaxial potential were also presented.
D. Vogt thanks FAPESP for financial support. P. S. Letelier thanks
CNPq and FAPESP for financial support. This research has made use of
NASA's Astrophysics Data System.
References
[1] W. Jaffe, Mon. Not. R. Astron. Soc. 202, 995 (1983).
[2] L. Hernquist, Astrophys. J. 356, 359 (1990).
[3] M. Miyamoto and R. Nagai, Publ. Astron. Soc. Japan 27, 533 (1975).
[4] C. Satoh, Publ. Astron. Soc. Japan 32, 41 (1980).
[5] T. de Zeeuw and D. Pfenniger, Mon. Not. R. Astron. Soc. 235, 949
(1988).
[6] K. Long and C. Murali, Astrophys. J. 397, 44 (1992).
[7] S. Binney and S. Tremaine, Galactic Dynamics, (Princeton University
Press, Princeton, 1987).
[8] T. Morgan and L. Morgan, Phys. Rev. 183, 1097 (1969).
[9] L. Morgan and T. Morgan, Phys. Rev. D 2, 2756 (1970).
[10] J. Bic´ak, D. Lynden-Bell and J. Katz, Phys. Rev. D 47, 4334 (1993).
[11] J. P. S. Lemos and P. S. Letelier, Phys. Rev. D 49, 5135 (1994).
[12] G. A. Gonz´alez and P. S. Letelier, Phys. Rev. D 62, 064025 (2000).
8
[13] G. A. Gonz´alez and P. S. Letelier, Phys. Rev. D 69, 044013 (2004).
[14] D. Vogt and P. S. Letelier, Phys. Rev. D 68, 084010 (2003).
[15] D. Vogt and P. S. Letelier, Phys. Rev. D 71, 084030 (2005).
[16] D. Vogt and P. S. Letelier, Mon. Not. R. Astron. Soc. 363, 268 (2005).
[17] R. Lutticke, R. J. Dettmar and M. Pohlen, Astron. & Astrophys. S.
145, 405 (2000).
[18] R. Lutticke, M. Pohlen and R. J. Dettmar, Astron. & Astrophys. 417,
527 (2004).
9
|
astro-ph/0608034 | 2 | 0608 | 2007-01-18T12:33:54 | The accelerating universe and a limiting curvature proposal | [
"astro-ph",
"gr-qc",
"hep-th"
] | We consider the hypothesis of a limiting minimal curvature in gravity as a way to construct a class of theories exhibiting late-time cosmic acceleration. Guided by the minimal curvature conjecture (MCC) we are naturally lead to a set of scalar tensor theories in which the scalar is non-minimally coupled both to gravity and to the matter Lagrangian. The model is compared to the Lambda Cold Dark Matter concordance model and to the observational data using the gold SNeIa sample of Riess et. al. (2004). An excellent fit to the data is achieved. We present a toy model designed to demonstrate that such a new, possibly fundamental, principle may be responsible for the recent period of cosmological acceleration. Observational constraints remain to be imposed on these models. | astro-ph | astro-ph | The Accelerating Universe and a Limiting
Curvature Proposal
Damien A. Easson
Centre for Particle Theory, Department of Mathematical Sciences, Durham
University, Science Laboratories, South Road, Durham, DH1 3LE, U. K.
E-mail: [email protected]
Abstract. We consider the hypothesis of a limiting minimal curvature in gravity
as a way to construct a class of theories exhibiting late-time cosmic acceleration.
Guided by the minimal curvature conjecture (MCC) we are naturally lead to a set
of scalar tensor theories in which the scalar is non-minimally coupled both to gravity
and to the matter Lagrangian. The model is compared to the Lambda Cold Dark
Matter concordance model and to the observational data using the "gold" SNeIa
sample of Riess et. al. (2004). An excellent fit to the data is achieved. We present a
toy model designed to demonstrate that such a new, possibly fundamental, principle
may be responsible for the recent period of cosmological acceleration. Observational
constraints remain to be imposed on these models.
PACS numbers: 98.80.-k, 04.50.+h, 95.36.+x
arXiv: astro-ph/0608034
DCPT-06/17
7
0
0
2
n
a
J
8
1
2
v
4
3
0
8
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
The Accelerating Universe and a Limiting Curvature Proposal
2
1. Introduction
One of the most profound discoveries of observational physics is that the universe is
accelerating in its expansion [1, 2, 3, 4, 5, 6]. There have been many attempts to explain
this late-time acceleration, for example, a pure cosmological constant, dark energy
associated with some new scalar field and modified gravitational theories, although
all current models require some level of fine-tuning and none are considered to be a
complete explanation ‡. Whatever is responsible for the current acceleration may arise
from some completely new physical principle. This is the possibility we consider in this
paper. Our goal is to construct a toy model that represents a late-time accelerating
Universe using a new, possibly fundamental, principle. As our guiding principle, we
hypothesize the existence of a minimal curvature scale in gravity.
In a Friedmann, Robertson-Walker (FRW) space-time, without cosmological
constant Λ and with only standard matter sources such as dust and radiation, the
universe will always decelerate as it expands. One way to avoid this is to add matter to
the system that violates the strong energy condition (SEC). In a cosmological context
this violation constitutes the addition of matter sources satisfying the equation of state
p ≤ −1/3ρ. A second possibility is to explicitly remove flat space-time as a solution to
the theory. In this case the vacuum of the theory, which is approached at late times
as the energy density in matter fields becomes more and more dilute, is not Minkowski
space-time, but instead an accelerating Universe [9, 10, 11]. To remove flat spacetime
as a solution we hypothesize the existence of a minimal curvature in our underlying
fundamental theory. The simplest example of this is, of course, to introduce a bare
cosmological constant into General Relativity. However, in principle there may exist
many other ways to achieve this result.
Indeed, it appears that many accelerating
cosmological models derived from modified gravity theories contain such a minimal
curvature [12, 13, 14].
The idea of a minimal curvature scale in gravity mirrors that of a maximal curvature
scale.
In the literature many authors have considered this possibility and used it
to remove the curvature singularities of General Relativity by bounding curvature
invariants from above at the level of the classical action [15]-[26].
In the case of
singularity removal, it is necessary to bound all curvature invariants in order to cover
all possible physical situations in which such a singularity may occur.
By contrast, in the case of a minimal curvature approached at late times in a
homogeneous, isotropic universe, symmetry implies that it is only necessary to bound
the Ricci scalar R from below. Hence, unlike in the case of a maximal curvature
hypothesis, we shall see that one may implement a minimal curvature by using a modified
Brans-Dicke theory where the Brans-Dicke field couples non-minimally to the matter
Lagrangian.
Within this context we demonstrate that the existence of the minimal curvature
(MC) produces a Universe that evolves from a matter dominated period to an
‡ For a recent reviews on the subject see [7, 8].
The Accelerating Universe and a Limiting Curvature Proposal
3
accelerating phase mimicking the Λ-Cold-Dark-Matter (ΛCDM) model. We emphasize
that the model presented here is only a toy construction of the late Universe. The
model is not intended to provide a consistent cosmology from the time of Big-Bang
Nucleosynthesis (BBN) until today.
It is unlikely that the precise model presented
here is compatible with solar system experiments and the tight constraints on the time
variation of Newton's constant. However, the model does provide an example of how
the postulated existence of a minimal curvature scale in gravity can provide a new
mechanism to generate cosmological acceleration of the late Universe. Furthermore,
the model may capture features of a possibly more fundamental theory that admits a
minimal curvature scale S.
In Section 2, we describe the minimal curvature construction, first by using a toy
example and then by using a class of modified Brans-Dicke theories. We solve the
equations of motion for this example and demonstrate how the Universe evolves from
a matter dominated phase to an accelerating period as the curvature approaches its
minimal value.
In Section 3, we compare the MC model with ΛCDM and to the
supernovae (SNeIa) "gold" sample of [27]. Finally, we comment on the possibility
of constructing more realistic models that satisfy the limiting curvature hypothesis
and offer our conclusions and speculations in Section 4.
In Appendix A, we provide
a detailed analysis of the vacuum MC theory. In Appendix B, we construct an Einstein
frame description of the vacuum theory and compare it to the MC vacuum.
2. The Minimal Curvature Construction
Our goal is to construct theories in which a certain physical quantity is bounded from
below. Before leaping directly into our model, it is instructive to consider an example
of how a similar effect may be achieved in a simpler theory - the bounding of velocities
from above in Special Relativity by the speed of light [20].
The Newtonian action for a free particle of mass m in motion is
S =Z dt
1
2
m x2 .
(1)
In this classical theory the velocity of the particle is without bound. Now let us implement
one of the fundamental consequences of Special Relativity: To ensure that the speed of
this particle is limited by the speed of light we introduce a field φ(t) which couples to
the quantity in the action that we want to bound ( x2) and has a "potential" U(φ). The
resulting action is
(2)
(3)
S = mZ dt (cid:20) 1
2
x2 + φ x2 − U(φ)(cid:21) .
The variational equation with respect to φ
x2 =
∂U
∂φ
,
S An interesting example of a minimal curvature scale occurs in a certain classical limit of quantum
gravity [13, 14].
The Accelerating Universe and a Limiting Curvature Proposal
4
ensures that x is bounded, provided ∂U/∂φ is bounded. Note the absence of a kinetic
term for ϕ in the action, and hence, the reason the word potential appears in quotes
above. In order to obtain the correct Newtonian limit for small x and small φ we take
U(φ) proportional to φ2. In the Newtonian limit the action (2) reduces to (1). A simple
potential satisfying the above asymptotics is
U(φ) =
2φ2
1 + 2φ
.
(4)
Integrating out φ yields (up to an irrelevant constant) the action for relativistic particle
motion:
SSR = mZ dt√1 − x2 .
The above model provides a powerful example of how a toy construction based on a
fundamental principle -- the existence of a universal "speed limit" -- can capture features
of a more fundamental theory.
We now use a similar construction to model the existence of a minimal curvature
(MC) scale in gravity. Because we are interested in late time cosmology, we need only be
concerned with bounding one curvature invariant, the Ricci scalar R. In direct analogy
with our example from Special Relativity, we introduce a scalar field ϕ that couples to
the quantity we wish to bound, R, and a "potential" function V (ϕ)
SM C =Z d4x√−g M 2
pl
2
R − γϕR + V (ϕ)! ,
(6)
where Mpl ≡ (8πG)−1/2 is the reduced Planck mass and both ϕ and γ posses dimensions
of mass, [ϕ] = [γ] = M. The vacuum theory is equivalent to a Brans-Dicke theory with
Brans-Dicke parameter ωBD = 0. This is seen by re-writing the action (6) in terms of a
new field
(5)
(7)
(8)
(9)
so that the action becomes
pl
2 − γϕ! ,
Φ = M 2
SBD =Z d4x√−g (ΦR + V (Φ)) .
Ordinarily, such a theory can be re-cast as a purely gravitational theory with
Lagrangian L = √−gf (R) (see, e.g. [29]); however, this is not possible for all forms
of the potential V .
For reasons that will become clear shortly, we allow the field ϕ to couple non-
minimally to matter. The non-minimal coupling yields the matter action,
SM =Z d4x√−g LM (ψi, ϕ, f (ϕ/Mpl) ¯ψψ) ,
where LM is the Lagrangian made up of whatever matter fields ψi are in the theory. In
the case were ψ represents a dark matter Dirac spinor, the field ϕ couples non-minimally
only to the dark matter sector and need not couple to baryons. In this case it is possible
to avoid constraints on such a coupling from solar-system and table-top tests of gravity.
The Accelerating Universe and a Limiting Curvature Proposal
5
In string theory, non-perturbative string loop effects do not generically lead to universal
couplings, allowing the possibility that the dilaton decouples more slowly from dark
matter than ordinary matter (see, e.g.
[30, 31, 32, 33, 34, 35, 36]).
This coupling can be used to address the coincidence problem, since the acceleration
is triggered by the coupling to matter. For the purposes of our toy construction, we do
not distinguish between baryonic and dark matter in the remainder of our discussion.
Note that in the case of f (R) theories which are conformally identical to models
of quintessence in which matter is coupled to dark energy with a large coupling,
this strong coupling induces a cosmological evolution radically different from standard
cosmology [37]. Similar difficulties may arise in the model presented here, however, we
have yet to investigate this issue.
The matter stress-energy tensor is given by
2
√−g
We assume a perfect fluid
Tµν ≡ −
δLM (ψi, ϕ)
δgµν
.
Tµν = (ρM + PM )uµuν + PM gµν ,
(10)
(11)
where uα is the fluid rest-frame four-velocity, and the energy density ρM and pressure
PM are related by the equation of state PM = wρM . Because we are focusing on the late
Universe we shall ignore the presence of radiation and consider only a matter density
ρM , which redshifts with expansion in the usual manner (with the exception of the
non-trivial ϕ dependence)
ρM ∝
f (ϕ/Mpl)
a3
.
(12)
In the above, f is a function describing the non-minimal coupling of the field ϕ to the
matter Lagrangian. Such couplings in the context of ordinary scalar-Einstein gravity
were studied in [36] where it was found that this coupling can be made consistent
with all known current observations, the tightest constraint coming from estimates of
the matter density at various redshifts. This coupling plays a critical role in Modified
Source Gravity, introduced in [38].
Variation of the total action Stot = SM C + SM with respect to the metric tensor,
δStot/δgµν = 0 yields the modified Einstein equations
(1 − αϕ)Rµν −
1
2
(1 − αϕ) R gµν
+ α∇µ∇νϕ − αgµν ✷ϕ − V gµν = 8πG Tµν
(13)
where ⊓⊔ ≡ gµν∇µ∇ν and we have introduced α ≡ 2γ/M 2
the field ϕ gives
pl. Variation with respect to
γR =
∂V
∂ϕ
+
1
f0
∂f (ϕ/Mpl)
∂ϕ
ρM .
(14)
Equation (14) is the key to imposing the limiting curvature construction. It is clear that
the curvature R will remain bounded and approach a constant curvature at late times
ds2 = −dt2 + a2(t)ndr2 + r2dΩ2o ,
(15)
The Accelerating Universe and a Limiting Curvature Proposal
6
(t → ∞) as long as V ′ + f ′ρM → constant; this is the essence of the construction. We
assume a flat (k = 0) Friedmann-Robertson-Walker metric
where a(t) is the scale factor of the Universe and dΩ is the line-element of the unit
2−sphere. Defining the Hubble parameter by H ≡ a/a and substituting the metric
ansatz (15) into (13) and (14) gives the generalized Friedmann equation (the 00-
component of (13)) and the equation of motion for ϕ:
3M 2
plH 2 − 6γϕH 2 − 6γH ϕ + V (ϕ) =
6γ(2H 2 + H) − V ′(ϕ) = f ′ ϕ
Mpl! ρ(0)
f0a3 ,
M
f (ϕ/Mpl)ρ(0)
M
f0a3
,
(16)
M
where a prime denotes differentiation with respect to ϕ, f0 ≡ f (ϕ0/Mpl), ϕ0 and ρ(0)
are the values of ϕ and ρM today.
By considering the asymptotics of our cosmology at early and late times, we can
constrain the forms of the functions V (ϕ) and f (ϕ). We require that the effective
Newton's constant for our theory remain positive definite so that gravity is always
attractive. This imposes a constraint on Φ ≥ 0, ∀ t in equation (8). There are
rather strong constrains on the time variation of Newton's constant from the period of
nucleosynthesis until today (roughly,(cid:12)(cid:12)(cid:12) G/G(cid:12)(cid:12)(cid:12) < O(10−10 − 10−13 yr−1) [39]). For the time
being, we will allow ourselves to ignore this constraint in order to produce a toy model
capable of realizing the MC conjecture. Furthermore, because we are only interested
in the behavior of the Universe from the matter dominated epoch until today, we have
ignored the presence of radiation. The absolute earliest our theory is valid is up to the
period of equal matter and radiation domination teq. For specificity, by early times we
refer to times near the time of photon decoupling at a redshift of z ≃ 1100, during which
the Universe is typically already well into the matter dominated regime.
At late times (and low curvatures) we want to bound the Ricci scalar R from below.
This will constitute a successful example of a model obeying the minimal curvature
hypothesis. To bound the curvature we use equation (14). It is clear that the curvature R
will remain bounded if we bound V ′+f ′ρM , where the prime denotes differentiation with
respect to ϕ. Hence, we require γ −1(V ′ + f ′ρM ) → R at late times, where we denote the
hypothesized minimal curvature scale by R. We anticipate that, by construction, there
will be a late-time attractor that is a constant curvature space-time with R = R. This
attractor is not an actual solution to the equations of motion. The above considerations
restrict the functional forms of potential V and the non-minimal coupling function f .
Integration singles out a class of theories that must obey V (ϕ) + f (ϕ)ρM → γRϕ as
t → ∞. The simplest forms for V and f obeying the above constraint are the linear
functions
V (ϕ) = µ3ϕ ,
f (ϕ) =
ϕ
Mpl
,
(17)
The Accelerating Universe and a Limiting Curvature Proposal
7
where µ is, in principle, another free parameter with dimensions of mass. However, we
will take µ = γ, so that R = γ2, and γ is the only free parameter in the theory k. We
now rescale time t ≡ tH0 so that today t0 = 1 and introduce the following dimensionless
quantities
a ≡
a
a0
ΩM ≡
, H ≡
ρM
ρc
=
ϕ
Mpl
H
H0
ϕ ≡
ρM
, R =
3H 2M 2
pl
, γ ≡
R
H 2
0
,
γ
Mpl
and take a0 = 1, where a0 is the value of the scale factor today.
In terms of the dimensionless quantities the EOM become
V (ϕ)
3
=
f (ϕ)Ω(0)
M
f0a3
,
H 2 − 2γϕH 2 − 2γH ϕ +
γ(2H 2 + H) −
V ′(ϕ)
6
= f ′ (ϕ)
Ω(0)
M
2f0a3 ,
(18)
(19)
(20)
(21)
The successful implementation of the minimal curvature hypothesis is now apparent.
Recasting equation (21) in terms of the curvature scalar R = 6(2H 2 + H), and
substituting in our choices for V and f (17):
R = γ2 + 3
Ω(0)
M
a3 .
(22)
We see that, as the Universe expands and the matter term dilutes, we asymptotically
approach the minimal value of the curvature R = R = γ2.
It is both interesting and surprising that the solution to Eq. (22) reduces to the
simple case of ΛCDM plus an arbitrary amount of dark radiation which may have either
positive or negative effective energy density. Most notably, this model arises in Randall-
Sundrum brane cosmology which has been extensively studied in the literature ¶. The
Friedmann equation derived from the Randall-Sundrum model for a flat Universe is
H 2 =
ρ
3M 2
pl
+
ρ2
36M 6
5
+ C
a4 ,
(23)
where M5 is the five dimensional Planck mass and Ca−4 is the so called dark radiation
term, since it scales like radiation, but it's origins are purely gravitational and it does
not interact with standard matter [45]. At low energies (when the energy density is
much less than the critical brane tension), the ρ2 term can be safely neglected. The
main observational restrictions on the dark radiation term come from the acoustic
scale at recombination (see, e.g. [46]), and from the amount of total growth of density
perturbations in the non-relativistic matter component from the time of equal matter
and radiation until the present day [47, 48]. As a result, the density of dark radiation
cannot be significantly larger than the present CMB energy density +.
k It is interesting to note that if we do not take µ = γ and then take the limit that γ → 0 the theory
resembles the action of the modified source gravity models studied in [38].
¶ For some reviews see, e.g. [43, 44].
+ Note, the constraints from BBN are even stronger [49].
The Accelerating Universe and a Limiting Curvature Proposal
8
Making use of (18) and (19), Eq. (23) may be recast (neglecting the ρ2 term) as
H 2 =
ΩM
a3 +
ΩR
a4 +
ΩΛ
3
,
(24)
where we have included a cosmological constant term ΩΛ = Λ/3M 2
plH 2, and ΩR =
ΩC + Ωr includes contributions from both the dark and ordinary radiation. From (24)
we find
H = −3ΩM
2a3 −
2ΩR
a4
.
(25)
Constructing the Ricci Scalar R = 6(2H 2 + H) from the above expressions yields
Eq. (22), with γ2 = 4ΩΛ.
In Appendix A, we provide a detailed analysis of the vacuum MC equations with
ΩM = 0. Although the presence of matter plays an important role in our minimal
curvature construction, an analysis of the vacuum theory provides valuable insight into
the solutions we are interested in studying. In Appendix B, we transform the vacuum
MC theory into an Einstein frame and relate quantities of physical interest in both
frames.
For general solutions to the equations of motion (20) and (21) with functions (17)
we solve for the Hubble parameter H(t) and scalar field ϕ(t). We plot the relevant
portion of the ϕ−H phase space in Fig. 1.
2.5
2.25
2
1.75
1.5
1.25
1
0.5
1
1.5
2
2.5
Figure 1. Phase space of ϕ vs. H. All solutions asymptotically approach a late-time
accelerating phase with constant H at the minimal curvature R denoted by the red
line.
To solve the equation we integrate from the past to today (H0 = 1) and then
from today into the future and patch the solutions together. In the plots we take the
conditions H0 = 1.0, γ = 3.15, ΩM = 0.25 and let the values of ϕ0 vary. At late times,
the solutions approach the constant H attractor when the Universe is well into the
accelerating epoch.
The Accelerating Universe and a Limiting Curvature Proposal
9
3. Comparison with ΛCDM
We now compare our model with ΛCDM. To do so, we must enter reasonable initial
conditions into our numerical study and solve the equations of motion (20), (21) together
with (17) and the equation a = Ha. Let us begin by considering the value of the minimal
curvature. Physically the minimal curvature corresponds to an emptying of the matter
in the Universe due to cosmological expansion. In our model the value of the minimal
curvature is approached asymptotically. Today, the value of the curvature is given by
R0 = 6(2H 2
0 + H0) ≈ 12H 2
0 .
(26)
Because we observe a significant amount a matter in the Universe today, we know that
we have not yet reached the minimal curvature. However, because we are accelerating,
we know that we are near the minimal curvature (i.e. the first and second terms on the
right hand side of equation (22) must be comparable). Hence, from equation (26), we
expect the value of the minimal curvature R to be close to but less than R < 12H 2
0 .
Therefore, in terms of our dimensionless quantities (18), R < 12 and the free parameter
in our theory γ < √12 ≈ 3.4641. For the solutions considered below, we choose a
value of γ = 3.15 meeting the above requirements and that follows the ΛCDM model
to our satisfaction for our toy construction (i.e. a value leading to a matter dominated
cosmology followed by a "jerk" near a redshift of z ≃ .5 into an accelerating phase).
In the action (6), the effective Newton's constant is GNef f is given by
(16πGNef f )−1 =
M 2
pl
2 − γϕ .
(27)
To ensure that the effective Newton's constant remains positive definite over the history
of the Universe ( GNef f > 0, ∀ t) we must have γϕ < .5. We are almost in a position to
compare our model both with ΛCDM, which has the Friedmann equation
H =vuut Ω(0)
M
a3 + (1 − Ω(0)
M ) ,
and with the observational data provided by the SNeIa gold sample. To make a
comparison with the observational data we require an understanding of the luminosity
distance formula in the context of modified gravity models.
An important consideration arises when using the formula for the luminosity
(28)
distance in theories of the form:
S =Z d4x√−g "A(ϕ)
R
2
+ B(ϕ)(∂µϕ)2 + V (ϕ) +Xi
Ci(ϕ)Li# ,
(29)
where the Li represent the different types of possible matter Lagrangians present. Such
a theory arises as the low-energy effective action for the massless modes of dilaton
gravity in string theory, and our model is an action of just this sort [41, 42]; albeit, with
an unusual choice for the functions A, B, V, C. As we have already discussed, these
theories typically lead to time variation in Newton's gravitational constant. The time
In
variation can affect the way one should compare the theory to observations [50].
The Accelerating Universe and a Limiting Curvature Proposal
10
particular, the time-evolution can alter the basic physics of supernovae. For example,
the time variation in GN leads to different values of the Chandrasekhar mass at
different epochs, and hence, a supernova's peak luminosity will vary depending on when
the supernova exploded. This makes treating the supernovae Ia as standard candles
difficult [51, 52, 53, 54, 55, 56]. Specifically, the peak luminosity of SNeIa is proportional
to the mass of nickel synthesized which is a fixed fraction of the Chandrasekhar mass
MCh ∼ G−3/2. Hence, the luminosity peak of SNeIa varies as L ∼ G−3/2 and the
corresponding absolute magnitude evolves as
M = M0 +
15
4
log
G(z)
G0
,
(30)
where the subscript zero indicates the local values of the quantities. Therefore, the
magnitude-redshift relation of SNeIa in modified gravity theories of the type given by
(29) is related to the luminosity distance via [53, 55]:
m(z) = M0 + 5 log dL(z) +
15
4
log
G(z)
G0
.
(31)
Even if gravitational physics is described by some theory other than General Relativity
the standard formula for the luminosity distance applies as long as one is considering a
metric theory of gravity [57]:
dL(z; H(z), H0) =
(1 + z)
H0
Z z
0
dz′
H(z′)
.
For ΛCDM, the Luminosity distance (32) can be written [40]:
dL(z; ΩM , ΩΛ, H0) =
c(1 + z)
H0qκ
S(cid:18)qκZ z
2(cid:19)
0 h(1 + z′)2(1 + ΩM z′) − z′(2 + z′)ΩΛi− 1
(32)
(33)
where, S(x) ≡ sin(x) and κ = 1 − Ωtot for Ωtot > 1 while S(x) ≡ sinh(x) with
κ = 1 − Ωtot for Ωtot < 1 while S(x) ≡ x and κ = 1, for Ωtot = 1. Here and throughout,
Ωtot = ΩM + ΩΛ.
3.1. Numerical Analysis
Given the above considerations we take the following values of parameters today:
a0 = 1.0, H0 = 1.0, ϕ0 = 0.1,
γ = 3.15, Ω(0)
M = 0.25 .
(34)
We then integrate our equations from the past to today and then from today into the
future and patch the solutions together. While we do not provide an exhaustive study of
the parameter space for the MC model, the parameters given above provide a successful
example of the construction which fulfills our rather modest goals ∗. It is quite possible
that the parameters (34) can be tuned to achieve an even better agreement with ΛCDM.
∗ An exhaustive study of the phase space of the vacuum theory is supplied in Appendix A.
The Accelerating Universe and a Limiting Curvature Proposal
11
3.2. Results
The history and future of the curvature R together with the co-moving Hubble radius
are plotted in Fig. 2.
R
15
14
13
12
11
10
CURVATURE
Comoving Hubble Radius
Log1aH
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
0
1
2
t
3
4
-2
-1.5
-1
Logt
-0.5
0
Figure 2. The Ricci scalar (left) and co-moving Hubble radius (right) as functions of
cosmic time t. The curvature decreases from the matter-dominate past (red curve) to
the constant minimal value at R = γ 2 indicated by the green line. Today, t0 = 1 and
the blue curve indicates the future evolution of the curvature. While the Universe is
accelerating the co-moving Hubble radius decreases.
In the Figure the past history of the curvature is plotted in red while the future is
plotted in blue. Today we are at the value t = 1. The minimal curvature is given by
the green line at R = R = γ2. As expected, the curvature is large in the past when
the Universe is matter dominated and then decreases, approaching the accelerating
late-time de-Sitter attractor with constant minimal curvature R = R. The second
plot in the Figure shows the evolution of the co-moving Hubble radius H −1/a. The
phenomenologically desired cosmological transition from matter domination to late-
time acceleration of the Universe is clearly indicated by the decreasing of the co-moving
Hubble radius, when
H −1
< 0 .
d
dt
a
(35)
In some modified gravity models it can be difficult to achieve this transition (see,
e.g. [58, 59]).
The majority of our results are presented in Figures 3 and 4. In Figure 3 we plot
the scale factor a as a function of cosmic time t, the Hubble parameter H as function
of redshift z, where
z + 1 ≡
a0
a
,
the deceleration parameter
a
aH 2 ,
q(z) = −
and the luminosity distance, dL.
(36)
(37)
The Accelerating Universe and a Limiting Curvature Proposal
12
a
15
12.5
10
7.5
5
2.5
0
a. Scale Factor
b. Hubble Parameter
2.5
2.25
2
1.75
1.5
1.25
1
H
0
1
2
t
3
4
0.250.50.75 1 1.251.5
z
qz
c. Deceleration Parameter
1
0.75
0.5
0.25
0
-0.25
-0.5
-0.75
0.250.50.7511.251.5
z
d_L
d. Luminosity Distance
3
2.5
2
1.5
1
0.5
0
0 0.250.50.75 1 1.251.5
z
Figure 3. A comparison of the cosmological evolution of the MC model with ΛCDM.
Quantities plotted in red are the MC model while those in blue are for ΛCDM. Today
we are at the values t = 1, z = 0. Plot a.) shows the time evolution of the scale factor.
Plot b.) shows the Hubble parameter as a function of redshift. Plot c.) compares
the deceleration parameters as a function of redshift. Plot d.) shows the luminosity
distance, dL, as a function of redshift. The MC model is clearly a good fit with ΛCDM.
For the particular set of parameters considered, the scale factor of our model
differs from ΛCDM most strongly in the far future, although the difference in the
expansion rates of the MC construction with ΛCDM is apparent in the plot of the Hubble
parameters at high redshifts. The transition from matter domination to acceleration (the
jerk) occurs at a slightly lower redshift than ΛCDM. There is only a slight difference
in dL that occurs at high redshifts (z > 1), although the difference is not significant
enough to distinguish our model from ΛCDM using only the current supernova data.
The Accelerating Universe and a Limiting Curvature Proposal
13
Figure 4. Comparison of the MC and various ΛCDM models with observational data.
The supernovae data points are plotted with error bars and the data is taken from [27].
The luminosity distance dL for the MC model is plotted by the dashed (blue) curve.
The various theoretical predictions for ΛCDM are represented by the solid curves and
were examined in [28].
In Figure 4, we compare the luminosity distance predicted by our model with several
versions of ΛCDM and with the observational data. The luminosity distances are plotted
out to a redshift of z = 2 (the highest redshift supernova data is from z = 1.76). The
theoretical predictions of the minimal curvature construction and ΛCDM are compared
with the "gold" supernovae sample of [27]. The particular choice of ΛCDM models
shown are from [28]. The luminosity distance for the minimal curvature construction
is denoted by the blue dashed line and fits the supernova data extremely well. The
luminosity distance dL is virtually indistinguishable from ΛCDM with Ω(0)
M = 0.25 and
Ω(0)
Λ = 0.75 (there is a small discrepancy at high redshift as can be seen from Figure 3).
4. Conclusions
We have shown that a period of late-time cosmic acceleration can follow directly
from a simple minimal curvature conjecture (MCC). The model fits the SNeIa data
exceptionally well. While the specific formulation considered here is only a toy
The Accelerating Universe and a Limiting Curvature Proposal
14
construction, unlikely to be compatible with constraints from solar system and table
top test of the equivalence principle, it may capture phenomenologically interesting
features of a more fundamental theory that admits a limiting minimal curvature.
Furthermore, the construction successfully demonstrates the possibility that a new
fundamental physical principle may ultimately be responsible for the recent period of
cosmological acceleration. It is possible that experimentally viable models based on the
minimal curvature conjecture exist. The search for such models within the context of
scalar-Gauss-Bonnet gravity is currently underway. Despite the tight theoretical and
experimental constraints on scalar-Gauss-Bonnet cosmologies [62, 63, 64, 65, 66], we
remain optimistic that an experimentally and theoretically viable model based on the
minimal curvature construction can be discovered.
Acknowledgments
It is a pleasure to thank R. Brandenberger, R. Gregory, V. Jejjala, I. Moss, R. Myers
and T. Underwood for helpful discussions. I am especially grateful to M. Trodden for
numerous useful discussions over the course of this work. This work is supported in part
by PPARC and by the EU 6th Framework Marie Curie Research and Training network
"UniverseNet" (MRTN-CT-2006-035863).
References
[1] A. G. Riess et al.
[Supernova Search Team Collaboration], Astron. J. 116, 1009 (1998);
[arXiv:astro-ph/9805201];
[2] S. Perlmutter et al. [Supernova Cosmology Project Collaboration], Astrophys. J. 517, 565 (1999);
[arXiv:astro-ph/9812133];
[3] C. B. Netterfield et al.
[arXiv:astro-ph/0104460];
[Boomerang Collaboration], Astrophys. J. 571,
604 (2002);
[4] N. W. Halverson et al., Astrophys. J. 568, 38 (2002). [arXiv:astro-ph/0104489].
[5] J. L. Tonry et al., Astrophys. J. 594, 1 (2003); [arXiv:astro-ph/0305008];
[6] C. L. Bennett et al., Astrophys. J. Suppl. 148, 1 (2003); [arXiv:astro-ph/0302207];
[7] S. Nojiri and S. D. Odintsov, arXiv:hep-th/0601213.
[8] E. J. Copeland, M. Sami and S. Tsujikawa,
Int. J. Mod. Phys. D 15, 1753 (2006)
[arXiv:hep-th/0603057].
[9] S. M. Carroll, V. Duvvuri, M. Trodden and M. S. Turner, Phys. Rev. D 70, 043528 (2004)
[arXiv:astro-ph/0306438].
[10] S. M. Carroll, A. De Felice, V. Duvvuri, D. A. Easson, M. Trodden and M. S. Turner, Phys. Rev.
D 71, 063513 (2005) [arXiv:astro-ph/0410031].
[11] D. A. Easson, Int. J. Mod. Phys. A 19, 5343 (2004) [arXiv:astro-ph/0411209].
[12] S. Nojiri and S. D. Odintsov, Mod. Phys. Lett. A 19, 627 (2004) [arXiv:hep-th/0310045].
[13] F. P. Schuller and M. N. R. Wohlfarth, Phys. Lett. B 612, 93 (2005) [arXiv:gr-qc/0411076].
[14] F. P. Schuller and M. N. R. Wohlfarth, Nucl. Phys. B 698, 319 (2004) [arXiv:hep-th/0403056].
[15] M. A. Markov, Pisma Zh. Eksp. Teor. Fiz. 36, 214 (1982); Pisma Zh. Eksp. Teor. Fiz. 46, 342
(1987); V. P. Frolov, M. A. Markov and V. F. Mukhanov, Phys. Lett. B 216, 272 (1989).
[16] V. P. Frolov, M. A. Markov and V. F. Mukhanov, Phys. Rev. D 41, 383 (1990).
[17] D. Morgan, Phys. Rev. D 43, 3144 (1991).
[18] V. Mukhanov and R. H. Brandenberger, Phys. Rev. Lett. 68, 1969 (1992);
[19] M. Trodden, V. F. Mukhanov and R. H. Brandenberger, Phys. Lett. B 316, 483 (1993);
[arXiv:hep-th/9305111];
[20] R. H. Brandenberger, V. Mukhanov and A. Sornborger, Phys. Rev. D 48, 1629 (1993);
[arXiv:gr-qc/9303001].
The Accelerating Universe and a Limiting Curvature Proposal
15
[21] R. Moessner and M. Trodden, Phys. Rev. D 51, 2801 (1995); [arXiv:gr-qc/9405004].
[22] R. H. Brandenberger, R. Easther and J. Maia, JHEP 9808, 007 (1998); [arXiv:gr-qc/9806111].
[23] D. A. Easson and R. H. Brandenberger, JHEP 9909, 003 (1999); [arXiv:hep-th/9905175].
[24] D. A. Easson and R. H. Brandenberger, JHEP 0106, 024 (2001); [arXiv:hep-th/0103019].
[25] D. A. Easson, JHEP 0302, 037 (2003); [arXiv:hep-th/0210016];
[26] D. A. Easson, Phys. Rev. D 68, 043514 (2003). [arXiv:hep-th/0304168].
[27] A. G. Riess et al. [Supernova Search Team Collaboration], Astrophys. J. 607, 665 (2004)
[arXiv:astro-ph/0402512].
[28] T. R. Choudhury
and T. Padmanabhan, Astron. Astrophys.
429,
807
(2005)
[arXiv:astro-ph/0311622].
[29] G. Magnano and L. M. Sokolowski, Phys. Rev. D 50, 5039 (1994) [arXiv:gr-qc/9312008].
[30] T. Damour, G. W. Gibbons and C. Gundlach, Phys. Rev. Lett. 64, 123 (1990).
[31] R. Bean and J. Magueijo, Phys. Lett. B 517, 177 (2001) [arXiv:astro-ph/0007199].
[32] R. Bean, Phys. Rev. D 64, 123516 (2001) [arXiv:astro-ph/0104464].
[33] T. Biswas and A. Mazumdar, arXiv:hep-th/0408026.
[34] T. Biswas, R. Brandenberger, A. Mazumdar and T. Multamaki, Phys. Rev. D 74, 063501 (2006)
[arXiv:hep-th/0507199].
[35] R. Bean, S. Carroll and M. Trodden, arXiv:astro-ph/0510059.
[36] S. Das, P. S. Corasaniti and J. Khoury, Phys. Rev. D 73, 083509 (2006) [arXiv:astro-ph/0510628].
[37] L. Amendola, D. Polarski and S. Tsujikawa, arXiv:astro-ph/0603703.
[38] S. M. Carroll, I. Sawicki, A. Silvestri and M. Trodden, arXiv:astro-ph/0607458.
[39] J. P. Uzan, Rev. Mod. Phys. 75, 403 (2003) [arXiv:hep-ph/0205340].
[40] S. Perlmutter et al. [Supernova Cosmology Project Collaboration], Astrophys. J. 483, 565 (1997)
[arXiv:astro-ph/9608192].
[41] T. Damour and A. M. Polyakov, Gen. Rel. Grav. 26, 1171 (1994) [arXiv:gr-qc/9411069].
[42] T. Damour and A. M. Polyakov, Nucl. Phys. B 423, 532 (1994) [arXiv:hep-th/9401069].
[43] D. A. Easson, Int. J. Mod. Phys. A 16, 4803 (2001) [arXiv:hep-th/0003086].
[44] D. Langlois, Prog. Theor. Phys. Suppl. 148, 181 (2003) [arXiv:hep-th/0209261].
[45] P. Binetruy, C. Deffayet, U. Ellwanger and D. Langlois, Phys. Lett. B 477, 285 (2000)
[arXiv:hep-th/9910219].
[46] Y. Wang and P. Mukherjee, Astrophys. J. 650, 1 (2006) [arXiv:astro-ph/0604051].
[47] M. Doran and G. Robbers, JCAP 0606, 026 (2006) [arXiv:astro-ph/0601544].
[48] E. V. Linder, Astropart. Phys. 26, 16 (2006) [arXiv:astro-ph/0603584].
[49] D. Langlois, Prog. Theor. Phys. Suppl. 163, 258 (2006) [arXiv:hep-th/0509231].
[50] V. Acquaviva, C. Baccigalupi, S. M. Leach, A. R. Liddle and F. Perrotta, Phys. Rev. D 71, 104025
(2005) [arXiv:astro-ph/0412052].
[51] L. Amendola, P. S. Corasaniti and F. Occhionero, arXiv:astro-ph/9907222.
[52] E. Garcia-Berro, E. Gaztanaga, J. Isern, O. Benvenuto and L. Althaus, arXiv:astro-ph/9907440.
[53] A. Riazuelo and J. P. Uzan, Phys. Rev. D 66, 023525 (2002) [arXiv:astro-ph/0107386].
[54] E. Gaztanaga, E. Garcia-Berro, J. Isern, E. Bravo and I. Dominguez, Phys. Rev. D 65, 023506
(2002) [arXiv:astro-ph/0109299].
[55] S. Nesseris and L. Perivolaropoulos, Phys. Rev. D 73, 103511 (2006) [arXiv:astro-ph/0602053].
[56] R. Gannouji, D. Polarski, A. Ranquet and A. A. Starobinsky, JCAP 0609, 016 (2006)
[arXiv:astro-ph/0606287].
[57] C. Shapiro and M. S. Turner, arXiv:astro-ph/0512586.
[58] D. A. Easson, F. P. Schuller, M. Trodden and M. N. R. Wohlfarth, Phys. Rev. D 72, 043504 (2005)
[arXiv:astro-ph/0506392].
[59] R. Punzi, F. P. Schuller and M. N. R. Wohlfarth, arXiv:gr-qc/0605017.
[60] H. K. Jassal, J. S. Bagla and T. Padmanabhan, Phys. Rev. D 72, 103503 (2005)
[arXiv:astro-ph/0506748].
[61] H. K. Jassal, J. S. Bagla and T. Padmanabhan, arXiv:astro-ph/0601389.
[62] S. Nojiri and S. D. Odintsov, Phys. Lett. B 631, 1 (2005) [arXiv:hep-th/0508049].
[63] G. Cognola, E. Elizalde, S. Nojiri, S. D. Odintsov and S. Zerbini, Phys. Rev. D 73, 084007 (2006)
[arXiv:hep-th/0601008].
[64] A. De Felice, M. Hindmarsh and M. Trodden, JCAP 0608, 005 (2006) [arXiv:astro-ph/0604154].
[65] G. Calcagni, B. de Carlos and A. De Felice, Nucl. Phys. B 752, 404 (2006) [arXiv:hep-th/0604201].
[66] T. Koivisto and D. F. Mota, Phys. Lett. B 644, 104 (2007) [arXiv:astro-ph/0606078].
The Accelerating Universe and a Limiting Curvature Proposal
16
Appendix A. Probing the Vacuum Theory
The basic mechanism used to implement the limiting curvature construction requires a
non-minimal coupling of ϕ to the matter Lagrangian. This coupling plays an important
role in the overall cosmological dynamics discussed above. Hence, an understanding of
the vacuum theory is perhaps not as fruitful as it would be in case of a theory with
minimal coupling of ϕ to matter; however, once the Universe becomes sufficiently large
and dilute the dynamics will resemble those of the vacuum theory. As we shall see the
vacuum MC theory is significantly richer than that of ΛCDM. The equations describing
the vacuum are given by the equations (16) with ΩM = 0. The dimensionless forms are
H 2 − 2γϕH 2 − 2γH ϕ +
γ(2H 2 + H) −
V ′(ϕ)
6
= 0 ,
V (ϕ)
3
= 0 ,
(A.1)
where we have re-introduced the parameter µ from equation (17) (previously set equal
to γ) for completeness and the potential is V (ϕ) = µ3ϕ. In the case of the vacuum we
may solve exactly for H(t). We focus on two types of solutions of particular interest.
In the first, the Hubble parameter is given by
whereR = µ3/γ is the minimal curvature and α is a constant. In this case
and the scale factor evolves as
,
3
(t − 6αγ)
H = R
6
12R tanh
H(t) =s 1
cosh
sR
a(t) = a0sech
sR
(t − 6αγ)
(6αγ − t)
sR
1
2
.
,
3
3
(A.2)
(A.3)
(A.4)
The quantities mentioned above, along with the field ϕ are plotted in Fig. (A1). In this
case the Universe undergoes a bounce as indicated in plots a.) and b.) in Fig. (A1).
After the bounce the Hubble parameter is pulled up to the value at the minimum
HdS =qR/12 ,
curvature H →qR/12. This is a de Sitter solution of the vacuum theory with
while ϕ continues to evolve exponentially ϕ(t)+(2γ)−1 ∝ exp(qR/12). Interestingly, H
and H evolve in such a way that the curvature R = 6(2H + H) is constant throughout
the evolution of the solution, fixed at the minimal value R = R. This behavior is not
surprising as the constraint on R comes from the second equation in the EOM (A.1).
(A.5)
The second family of solutions are of greater relevance. They are the vacuum
analogs of the solutions plotted in Fig. (1) and discussed at the end of Section 2. These
solution are pulled down to the value of H at the minimal curvature. The entire phase
space of solutions to the vacuum theory is shown in Fig. (A2). Representative solutions
The Accelerating Universe and a Limiting Curvature Proposal
17
a. Hubble Parameter
b.
0.3
0.2
0.1
H
0
-0.1
-0.2
-0.3
4
3.5
3
a
2.5
2
1.5
dHdt
0.15
0.125
0.1
0.075
0.05
0.025
0
3
2.5
2
phi
1.5
1
0.5
0
0
8
10
12
0
2
4
6
t
d.
2
4
6
t
8
10
12
0
2
4
6
t
8
10
12
c. Scale Factor
1
0
2
4
8
10
12
6
t
Figure A1. Various quantities of interested plotted as functions of cosmic time t. In
the plots we take R = γ = µ = α = 1.
of the two families of solutions discussed above are plotted by the solid curves. The
red curve is the solution given by equation (A.2). The second set of solutions (relevant
to Section 2) are plotted in green in the upper right quadrant of the phase space.
The system defined by the vacuum equations (equation (A.1)) has two unstable saddle
equilibrium points at the values
(A.6)
(ϕ, H) =−
M 2
pl
2γ
12 .
, ±sR
The equilibrium points are marked by the purple dots in Fig. (A2).
The Accelerating Universe and a Limiting Curvature Proposal
18
2
1.5
1
0.5
0
-0.5
-1
-1.5
-1.5
-1
-0.5
0
0.5
1
1.5
2
Figure A2. The phase space of the vacuum theory. H is plotted on the vertical axis
and ϕ is plotted on the horizontal axis. The vector field for solutions is drawn in black.
Solutions of particular interests are plotted by the solid curves. The constant de Sitter
value of H at the minimal curvature is indicated by the dashed (blue) line. The two
unstable saddle equilibrium points are designated by the purple dots. In the plot we
take γ = µ = 3.15.
Appendix B. An Einstein Frame Description
It is not our intent to provide a complete Einstein frame analysis of the full MC system
with matter; however, it is instructive to consider the vacuum theory in an Einstein
frame. To move to the Einstein frame we begin with the Brans-Dicke frame defined by
the action (8). Passage to the Einstein frame is achieved via a conformal transformation
of the form
gµν = Ω2gµν ,
(B.1)
where Ω2 is the conformal factor which must be positive to leave the signature of the
metric unaltered. From this point forward a tilde shall denote a quantity built out of
the Einstein-frame metric tensor gµν. Under this transformation the infinitesimal line
The Accelerating Universe and a Limiting Curvature Proposal
element and the determinant of the metric transform as
respectively. The Ricci scalar transforms as
Omitting the ordinary divergence √−ge⊓⊔(ln Ω), the action in the Brans-Dicke frame
transforms to
! .
ds2 = Ω2ds2 ,
q−g = Ω4√−g ,
R = Ω2 R + 6e⊓⊔(ln Ω) − 6gµνf∇µΩf∇νΩ
S =Z d4xq−g Φ
γ M 2
R − 6
2 − Φ! .
V (Φ) =
Ω2
Ω2
µ3
Φ
pl
Ω4 gµνf∇µΩf∇νΩ +
V (Φ)
Ω4 ! ,
where the potential in terms of Φ is given by (7) together with (17):
Here we have re-introduced the parameter µ from equation (17) (previously set equal
to γ) for completeness.
By choosing our conformal factor to be Ω2 = 2Φ/M 2
pl, and performing a field
redefinition to define the Einstein frame field φ,
19
(B.2)
(B.3)
(B.4)
(B.5)
(B.6)
(B.7)
(B.8)
(B.9)
the action (B.4) becomes the Einstein frame action
where we have defined the potential
R −
1
2
gµνf∇µφf∇νφ − eV (φ)! ,
,
3
φ
Φ =
pl
2
M 2
pl
2
s 2
exp
Mpl
SEF =Z d4xq−g M 2
eV (Φ) = − M 2
2 !2
eV (φ) =
M 2
pl
Φ2
pl
V (Φ)
.
The Einstein frame potential is given by
2 R(cid:16)eβφ/Mpl − 1(cid:17) e−2βφ/Mpl ,
where R = µ3/γ is the minimal curvature and β =q2/3.
In this frame the field φ has a canonically normalized kinetic term making the
interpretation of solutions more simple due to our familiarity with minimally coupled
scalar field to ordinary Einstein gravity. It is important to note, however, that so far
we have only considered the vacuum theory. Even in the Einstein frame the Einstein
field φ will couple non-minimally to matter and therefore, when matter is present in
significant amounts, the simple Einstein frame vacuum solutions will not be an accurate
description of the theory.
Under the conformal transformation (B.1), the cosmic time coordinate transforms
as
dt2 = eβφ/Mpl dt2 .
(B.10)
The Accelerating Universe and a Limiting Curvature Proposal
Taking the Einstein-frame Friedmann-Robertson-Walker flat metric
ds2 = −dt2 + a2(t)dx2 ,
Leads to the familiar equations of motion
H 2 =
and
1
pl (cid:18)1
2
φ′2 + V (φ)(cid:19) ,
3M 2
φ′′ + 3 Hφ′ +
d V
dφ
= 0 ,
20
(B.11)
(B.12)
(B.13)
where prime denotes differentiation with respect to the Einstein frame cosmic time
coordinate t and H ≡ a′/a. Using the conformal transformation (B.1), we find the
Hubble parameters in the Einstein and Minimal Curvature frames are related by
(B.14)
H = e
βφ
2Mpl H −
β
2Mpl
φ′! .
We plot the Einstein frame potential (B.9) in Fig. (B1). From the plot it is clear
0.2
0.1
0
-0.1
-0.2
-0.3
0
2
4
6
8
10
Figure B1. The Einstein frame potential eV as a function of the Einstein frame field
φ. In the plot we have set 16πGN = γ = µ = 1.
that there are three regions of interest:
i) There is an unstable de Sitter solution corresponding to the field sitting on the
top of the potential.
ii) If the field starts initially to the left of the unstable de Sitter solution (and
without significant positive velocity) the system will quickly run off to large negative φ.
iii) If the field falls to the right of the de Sitter solution it will asymptotically roll
to large positive values.
The Accelerating Universe and a Limiting Curvature Proposal
21
Let us examine the first possibility in greater detail. The unstable de Sitter solution
with constant H = HdS sits at the value at the maximum of the potential (B.9). The
maximum is at constant value
φ = φdS = β −1Mpl ln 2 ,
From equation (B.12) we find the corresponding H
HdS =
1
2sR
6
.
(B.15)
(B.16)
From equation (B.14) we see that the corresponding value in the minimal curvature
frame is
H( HdS, φdS) =sR
12
,
(B.17)
which is the exact de Sitter solution we found in our analysis of the vacuum of the
minimal curvature theory (A.5).
The Minimal Curvature frame field is related to the Einstein frame field via
ϕ(φ) =
and consequently,
φ(ϕ) =
M 2
pl
2γ (cid:16)1 − eβφ/Mpl(cid:17) ,
ϕ! .
ln 1 −
2γ
M 2
pl
Mpl
β
Using the relation (B.18), we find the value of ϕ at the de Sitter point (B.15):
ϕ(φdS) = −
M 2
pl
2γ
.
(B.18)
(B.19)
(B.20)
Hence we conclude that the unstable de Sitter solution in the Einstein frame is mapped
to one of the unstable saddle critical points in the vacuum of the Minimal Curvature
frame (see equation (A.6)).
We now examine case iii): when φ rolls to large positive values the Einstein frame
potential may be approximated by
leading to the exact solutions
M 2
plR
2
eV (φ) ≃
= √6 lnseV0
φ(t)
Mpl
24
H(t) =
3
t
,
e−βφ/Mpl ,
t
Mpl
(B.21)
(B.22)
(B.23)
where fV0 = M 2
plR/2, corresponding to power-law acceleration in the Einstein frame with
a(t) = a0t3 .
(B.24)
To find the corresponding behavior in the Minimimal Curvature frame we use the
conformal transformation (B.1) along with the transformation for the cosmic time
The Accelerating Universe and a Limiting Curvature Proposal
22
coordinate (B.10). We find the late-time power law attractor in the Einstein frame
(B.24) is mapped to the asymptotic de Sitter attractor in the MC frame at the minimal
curvature R (depicted by the dashed blue line in Fig. (A2)),
a(t) = a0e√ R
12 t .
(B.25)
|
astro-ph/9503016 | 1 | 9503 | 1995-03-03T11:45:34 | The Orion OB1 Association II. The Orion-Eridanus Bubble | [
"astro-ph"
] | Observations of the interstellar medium in the vicinity of the Orion OB1 association show a cavity filled with hot ionized gas, surrounded by an expanding shell of neutral hydrogen (the Orion-Eridanus Bubble). In this paper we examine this cavity and the surrounding bubble with the aid of data from the Leiden/Dwingeloo HI survey. We present neutral-hydrogen maps for the Orion-Eridanus region which allow identification of the HI filaments and arcs delineating the Bubble and derivation of its expansion velocity. The HI data are compared to X-ray, CO and IRAS 100 micron data. Using models of wind blown bubbles that take the density stratification of the Galactic HI layer into account we show that the stellar winds and supernovae from stars in Orion OB1 can account for the size as well as the expansion velocity of the HI shell. However, density inhomogeneities in the ambient interstellar medium cause significant discrepancies between our model and the observed shell. | astro-ph | astro-ph | A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
10.15.1; 10.15.2 Orion OB1; 09.02.1; 09.09.1 Orion-Eridanus Bubble; 09.11.1; 09.19.1
The Orion OB1 association
II. The Orion-Eridanus Bubble
A.G.A. Brown1, D. Hartmann1; ? and W.B. Burton1
1 Sterrewacht Leiden, P.O. Box 9513, 2300 RA, Leiden, The Netherlands
Received: : : , accepted: : :
ASTRONOMY
AND
ASTROPHYSICS
3.3.1995
5
9
9
1
r
a
M
3
6
1
0
3
0
5
9
/
h
p
-
o
r
t
s
a
Abstract. Observations of the interstellar medium in the vicin-
ity of the Orion OB1 association show a cavity filled with
hot ionized gas, surrounded by an expanding shell of neu-
tral hydrogen (the Orion-Eridanus Bubble). In this paper we
examine this cavity and the surrounding bubble with the aid
of data from the Leiden/Dwingeloo H i survey. We present
neutral-hydrogen maps for the Orion-Eridanus region which
allow identification of the H i filaments and arcs delineating
the Bubble and derivation of its expansion velocity. The X-ray
emission from the Orion-Eridanus region is enhanced with re-
spect to the Galactic background emission. Comparison with
the H i data shows a detailed anti-correlation of the X-ray en-
hancement with kinematically-narrow features. This confirms
the association of the X-ray enhancement with the cavity in
H i. Comparison of the derived column densities in H i with the
IRAS 100 (cid:22)m intensities shows that the H i shell emission is
optically thin. This justifies a derivation of the mass of the H i
shell by direct conversion of the observed H i emission to col-
umn densities. Models of wind-blown bubbles are considered,
to investigate if the cavity was formed by the stellar winds and
supernovae from Orion OB1. Using a model that takes the den-
sity stratification of the Galactic H i layer into account, we show
that the stellar winds and supernovae from stars in Orion OB1
can account for the size as well as for the expansion velocity of
the H i shell. However, density inhomogeneities in the ambient
interstellar medium cause significant discrepancies between our
model and the observed shell.
Key words: open clusters and associations: general, individual:
Orion OB1 -- ISM: bubbles, individual objects: Orion-Eridanus
Bubble, kinematics and dynamics, structure
Send offprint requests to: A.G.A Brown
? Present address: Harvard-Smithsonian Center for Astro-
physics, Mail Stop 47, 60 Garden Street, Cambridge, MA
02138, USA
1. Introduction
OB associations play an important role in the propagation of
star formation and in the evolution of the interstellar medium,
in general marshalling various aspects of the structure and the
kinematics of the medium. A prototypical example of the mar-
shalling of the ISM by an association can be found around
Orion OB1. This stellar association consists of four Subgroups
(1a -- 1d, see Blaauw 1964) and is accompanied by several large-
scale features in the interstellar medium. These features include
the Orion Molecular Cloud (OMC) complex (see e.g. Kutner
et al. 1977 and Maddalena et al. 1986); Barnard's Loop (see
e.g. Goudis 1982); H(cid:11) emission structures which extend all the
way to the Eridanus constellation, some 50(cid:14) below the Galactic
plane (Sivan 1974; Reynolds & Ogden 1979, henceforth RO);
and a cavity in the neutral hydrogen distribution surrounded by
H i arcs showing expansion motions (Heiles 1976; Green 1991;
Green & Padman 1993). The diffuse Galactic H(cid:11) background
intensity is enhanced in this area (e.g., Reynolds et al. 1974), and
the same is true for the X-ray intensity (e.g., Burrows et al. 1993,
henceforth BSNGG). Cowie et al. (1979) measured ultraviolet
absorption lines of various ionization stages of C, N, Si, and
S for stars in or near Orion OB1 and detected a high-velocity,
low-ionization shell surrounding all of the Orion area ("Orion's
Cloak"), as well as a dense, clumpy, low-ionization shell ex-
panding at smaller velocity. The cavity in the H i emission also
shows up in the IRAS sky flux maps of the Orion-Eridanus area
shown by BSNGG.
The various observed features were interpreted by RO as
being consistent with a cavity of warm ionized gas surrounded
by an expanding shell of H i. The ionization of the cavity would
be maintained by the UV radiation from the early-type stars in
Orion OB1. They explained the existence of the shell as due to a
series of supernova explosions about 2 (cid:2) 106 years ago. Cowie
et al. (1979) interpreted the high-velocity shell in terms of the
radiative shock of a supernova which occurred some 3 (cid:2) 105
years ago, and the high H i column densities in terms of the
remnants of older supernova events which occurred 2 -- 4 (cid:2) 106
years ago.
2
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
More recently, BSNGG presented a model in which part of
the observed X-ray enhancement is due to hot gas filling a cavity
created by the stellar winds from early type stars in Orion OB1.
Both RO and BSNGG present schematic views of the Orion-
Eridanus Bubble in which it expands towards the Sun. In the
model of BSNGG the Orion-Eridanus Bubble interacts with the
Local Hot Bubble in which the Sun is evidently immersed.
In this paper we present a more detailed look at the neutral
hydrogen content of the Orion-Eridanus region, using the Lei-
den/Dwingeloo H i survey. This H i survey has better spatial-
and velocity-resolution than the surveys used previously for
studies of the Orion-Eridanus Bubble. To model the size and
expansion velocity of the Bubble we use detailed information
on the stellar content of Orion OB1 as described by Brown et al.
(1994, henceforth Paper I). We employ models which take into
account the density stratification of the Galactic H i layer.
This Paper is organized as follows. In Sect. 2 we give a
short description of the Leiden/Dwingeloo H i survey and the
characteristics of the data. In Sect. 3 we present the H i data
for the region of interest. A comparison with data at other
wavelengths is made in Sect. 4, in particular with the BSNGG
X-ray material. In Sect. 5 we derive column densities of the
H i filaments and the mass of the neutral hydrogen shell and
discuss the energetics of the Bubble and its relation to the stars
in Orion OB1. Section 6 contains the conclusions.
2. H i survey data in the region
The H i data used in this paper are from the recently completed
Galactic survey made with the 25 m Dwingeloo telescope1
(Hartmann 1994 and Hartmann & Burton 1995). The entire
sky north of declination (cid:0)30(cid:14) was mapped in the 21 cm emis-
sion line, using a helium-cooled FET amplifier receiver with
a system temperature of about 35 K, together with the 1024-
channel prototype of the Dwingeloo Auto-correlation Spec-
trometer (DAS). The telescope beam width is 36 ; the obser-
vations were sampled on a 0(cid:14). 5 lattice typically with 180 s
integration time. The data were obtained in total-power mode.
Frequency-switched bandpass calibrations were frequently ob-
served at one full bandpass (5 MHz) higher frequency. Af-
ter data reduction, the velocity coverage of the survey is
(cid:0)450 (cid:20) vLSR (cid:20) +400 km s(cid:0)1 at 1:03 km s(cid:0)1 resolution. (All ve-
locities quoted here are expressed relative to the Local Standard
of Rest. ) The mean sensitivity of the data is about 0.07 K.
The data were corrected for stray-radiation contamination
using the method of Kalberla (1978; see also Kalberla et al.
1980). His algorithm, originally developed for the Effelsberg
100 m telescope antenna pattern and using the Berkeley H i
survey (Weaver & Williams 1973; Heiles & Habing 1974) as
input, was adapted to calculate the stray radiation contribution
for all spectra in the current survey. This led to convolving the
measured Leiden/Dwingeloo input sky with the 25 m's antenna
1 The Dwingeloo 25 m telescope is operated by the Nether-
lands Foundation for Research in Astronomy (NFRA).
pattern. Details of this procedure are given in Hartmann (1994)
and Hartmann & Burton (1995). The calculated stray radiation
was subtracted from the data and a third-order polynomial base-
line was removed using an unbiased automated algorithm. A
low-amplitude (typically (cid:24) 0:008 K) standing wave was sub-
tracted, and sharp interference spikes were detected, modeled,
and removed.
The reduced data were re-sampled onto a common grid
(0(cid:14). 5 (cid:2) 0(cid:14). 5) to create a homogeneous data cube of dimensions
(`; b; v) = (721; 361; 1024) representing the entire sky. The ob-
served sky at (cid:14) > (cid:0)30(cid:14) fills about 70% of the data cube. The
H i data discussed here form a subset of the entire H i data cube,
(cid:20) b (cid:20) 0(cid:14). These lim-
within the limits 160(cid:14)
its encompass the Orion-Eridanus region as well as the Taurus
and Perseus constellations.
(cid:20) ` (cid:20) 240(cid:14)
; (cid:0)60(cid:14)
3. H i loops, filaments and shell fragments in Orion-
Eridanus
The Orion-Eridanus Bubble was identified earlier in H i by
Heiles (1976) and Heiles & Jenkins (1976). Much of the H i
structure that we discuss below is also visible in maps pre-
sented by Heiles (1984). Figure 1 shows a collection of ve-
locity slices through our H i subset, each of width one km s(cid:0)1,
between (cid:0)40 (cid:20) vLSR (cid:20) +40 km s(cid:0)1. The central velocity is
indicated in each frame. The various structures must be traced
in both spatial- and velocity-space, as is generally the case in
H i analyses.
Figures 2 through 6 show individually-scaled maps of the
Tb(cid:1)v(K km s(cid:0)1) for the indicated
column-density measure P
velocity intervals. These intervals were chosen after visually
identifying filaments and partial loops in Fig. 1. Adjacent to the
more-conventionally displayed H i data, Figs. 2 -- 6 also show an
unsharp-masked version of the maps. In this process (originally
applied to photographic plates to enhance the dynamic range
of low-amplitude fluctuations on high-intensity backgrounds
(cid:2) 10(cid:14) boxcar-smoothed
by Malin, 1979), we subtracted a 10(cid:14)
copy of the image from the original. The size of the smoothing
kernel roughly corresponds to the size of the intensity variations
of the large-scale structures. The resulting maps bring out the
filamentary structures beautifully. The filaments and loops that
we identify as part of the expanding shell are shown isolated in
Fig. 8. The maps in this figure are bitmaps showing the features
for the same velocity intervals as maps A -- E.
For orientation purposes Fig. 7 shows the location of the
Orion OB1 association with respect to the H i shell. The dots
in this figure show the positions of the brightest stars in the
Orion constellation and the contours show the outlines of the
Orion A and B molecular clouds as observed at 100 (cid:22)m (IRAS).
The ring around (`; b) = (195(cid:14)
; (cid:0)12(cid:14)) is the (cid:21)-Orionis ring (see
e.g., Wade 1957). The locations of the three main subgroups of
Orion OB1 are indicated by circles.
Map A ((cid:0)40; (cid:0)30) km s(cid:0)1 : The emission in this extreme ve-
locity interval is generally weak in regions away from the Galac-
tic plane, but in the unsharp-masked map one can nevertheless
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
3
trace a number of filaments forming a ring-like structure. The
filaments are located between b = (cid:0)50(cid:14) and (cid:0)30(cid:14) and ` = 175(cid:14)
and 210(cid:14). These structures correspond to the approaching side
of an expanding H i shell. At less extreme velocities, they merge
with the features in Map B. The structures seen in this map can
also be followed in Fig. 1, down to vLSR (cid:25) (cid:0)40 km s(cid:0)1.
Map B ((cid:0)28; (cid:0)3) km s(cid:0)1 : A gap in the H i at negative ve-
locities is clearly visible between latitudes (cid:0)50(cid:14) and (cid:0)20(cid:14)
and longitudes 170(cid:14) and 210(cid:14). The filament at ` = 210(cid:14) was
used by BSNGG to derive the mass of the expanding H i
shell. The H i gap has an irregular structure due to emission
at (`; b) = (190(cid:14)
; (cid:0)30(cid:14)). We suggest in Sect. 3 that this emis-
sion is probably associated with foreground material.
Map C ((cid:0)1; +8) km s(cid:0)1 : The largest coherent H i structure in
the Orion-Eridanus region can be seen in this map. It is the
closed loop running between ` = 170(cid:14) and 230(cid:14) and b = (cid:0)55(cid:14)
and (cid:0)17(cid:14). This structure is outlined well in the unsharp-masked
image, with the filament at b = 50(cid:14) especially prominent. The
cavity in H i surrounded by the loop structure shows emission
which can be attributed (partly) to foreground material (see
Sect. 3). The angular dimension of the loop perpendicular to
the Galactic plane is about 33(cid:14) and along the Galactic plane
its maximum dimension is about 38(cid:14). If we regard the loop as
roughly circular, then its radius is about 18(cid:14). If this H i shell is
at the distance of the Orion OB1 association derived in Paper
I as 380 pc, this implies a linear radius of about 120 pc. Also
visible in this map is H i associated with the molecular cloud
complex near the Orion OB1 association. The Orion A and B
clouds are located near (`; b) = (208(cid:14)
; (cid:0)16(cid:14)). The (cid:21)-Orionis
ring is visible surrounding (`; b) = (195(cid:14)
; (cid:0)12(cid:14)).
Map D (+9; +14) km s(cid:0)1 : The most prominent feature at the
velocities in this map is the filament located around b = (cid:0)50(cid:14).
It is a continuation of the loop structure identified in Map C.
The loop is now seen running between ` = 170(cid:14) and 220(cid:14)
and b = (cid:0)55(cid:14) and (cid:0)20(cid:14). It has decreased in size at these
velocities compared to those entering Map C in accordance
with the expectations for an expanding shell.
Map E (+15; +40) km s(cid:0)1 : The most noticeable feature in this
map is the almost-closed elliptical structure located between
` = 170(cid:14) and 215(cid:14) and latitudes b = (cid:0)53(cid:14) and (cid:0)30(cid:14). This
ellipse can be followed in Fig. 1 up to vLSR (cid:25) +40 km s(cid:0)1, and
represents the receding side of the H i shell.
From detailed inspections of these maps we conclude that
H i features associated with the Orion-Eridanus Bubble can be
followed in velocity space from vLSR = (cid:0)40 km s(cid:0)1 to vLSR =
+40 km s(cid:0)1.
4. Data at other wavelengths
In this section we compare the H i data with published data at
other wavelengths.
4.1. H(cid:11) data
An extensive study of the H(cid:11) emission in the Orion-Eridanus
area was made by RO. They studied the emission line profiles
and constructed a map of diffuse H(cid:11) emission which shows two
prominent H(cid:11) features, namely Barnard's Loop and a filament
running along right ascension 4h (from (`; b) = (181(cid:14)
; (cid:0)31(cid:14))
to (`; b) = (196(cid:14)
; (cid:0)40(cid:14))). Their map reveals that the diffuse
Galactic H(cid:11) background intensity is enhanced in the area of the
H i cavity. They furthermore derived an expansion velocity of
about 15 km s(cid:0)1 for the ionized gas, based on the observed line
splitting in the emission profiles. The largest approaching and
receding velocities they observed were (cid:0)30 and +26 km s(cid:0)1,
respectively. They derived a temperature of 8000 (cid:6) 2000 K
for the ionized gas inside the cavity. The ultraviolet radiation
from the O stars in Orion OB1 was regarded as the source
of ionization. As a model for the origin of the Bubble they
proposed the occurrence of a supernova, or series of supernovae,
about 2 (cid:2) 106 years ago, with a total energy of about 3 (cid:2) 1052
ergs. The Orion OB1 association is itself the most probable
location for the suggested supernovae.
4.2. X-ray data
The cavity in the neutral hydrogen distribution in Orion-
Eridanus also coincides with an enhancement in the background
X-ray intensity. The most comprehensive study of the X-ray
data in this region was carried out by BSNGG. They combined
their X-ray data with neutral hydrogen data from the Bell Labs
21 cm survey (Stark et al. 1992) and with dust data from the
IRAS mission. Maps were presented of the X-ray emission
in two energy bands (0:25 keV and 0:6 keV). Two compo-
nents in the morphology of the X-ray data were distinguished
by BSNGG. There is a hook-shaped feature, corresponding to
the morphology of the cavity in the H i emission apparent in
Map D, which BSNGG referred to as EXE1 (located between
b = (cid:0)50(cid:14) and b = (cid:0)30(cid:14) and ` = 180(cid:14) and ` = 210(cid:14)), and a
circular feature, corresponding to the H i cavity seen in Map
C at (`; b) = (210(cid:14)
; (cid:0)43(cid:14)), referred to as EXE2. The circular
feature only shows up in the low-energy X-ray data. They de-
rived temperatures of the X-ray emitting gas of 2:1 (cid:2) 106 and
1:6 (cid:2) 106 K for EXE1 and EXE2, respectively. Both X-ray
features are associated with distinct H i shells, where EXE1
is associated with the shell that surrounds the Orion-Eridanus
Bubble. The column density and mass of this shell were derived
from the filament at ` = 210(cid:14), visible in Map B, and from the
approaching cap, at negative velocities. The mass derived from
the approaching cap is an order of magnitude smaller than that
derived from the filament.
A model for the Orion-Eridanus Bubble was proposed by
BSNGG in which EXE1 is caused by a stellar-wind-blown cav-
ity, and in which EXE2 is a separate bubble of hot gas possibly
due to the wind from a single star (see their Fig. 9). In their
model the Orion-Eridanus Bubble interacts with local clouds,
roughly 100 pc from the Sun. Evidence for this interaction
comes from the low-energy X-rays which appear to be partly
4
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
absorbed by a filament of gas corresponding to the H(cid:11) filament
found by RO running along (cid:11) = 4h. Furthermore, a map of
the ratio of X-ray emission in the high-energy band to the total
emission shows evidence for absorption of low-energy X-rays
by gas associated with the cloud Lynds Dark Nebula (LDN)
1569 at (`; b) = (189(cid:14)
; (cid:0)36(cid:14)). To explain the discrepancy in
the column densities derived from the filament at ` = 210(cid:14) and
the cap, BSNGG suggest that the Bubble is elongated along the
line of sight and expands mostly towards the Sun (away from
the Orion A and B clouds).
Figures 9a -- d compare contour plots of the BSNGG X-
ray data with gray-scale images of H i emission. The veloc-
ity intervals for which the H i data are shown were chosen
to bring out the anti-correlation of the X-ray emission with
kinematically-narrow features in the H i emission. The velocity
intervals are (cid:0)1:0 km s(cid:0)1 to +4:0 km s(cid:0)1 in Figs. 9a and 9b,
and +13:5 km s(cid:0)1 to +18:5 km s(cid:0)1 in Figs. 9c and 9d. These
velocity intervals correspond roughly to those in Map C and
Map D. EXE2 was found by BSNGG to be associated with
a separate H i shell surrounding the circular cavity in Map C.
Indeed, Fig. 9a shows the occurrence of the low energy X-ray
emission within the circular H i cavity in Map C. However, Figs.
9c and 9d show that the X-ray emission from both energy bands
corresponds most closely to the gap in H i seen in Map D. This
suggests that the X-ray-emitting gas is associated with a single
coherent expanding H i shell. Based on the facts presented in
Sect. 4.3, we argue that the appearance of a cavity in Map C
around EXE2 may be caused by emission from foreground H i,
partly filling in the H i cavity associated with the Bubble. If
indeed the X-ray emission from both energy bands corresponds
to a single cavity in H i, the nature of EXE2 remains to be
investigated.
The greater detail available in the Leiden/Dwingeloo data
(compared to the Bell Labs data utilized by BSNGG) reveals
additional features in the H i morphology that anti-correlate
with the X-ray emission. Figures 9c and 9d show the X-ray-
absorbing H i gas associated with LDN 1569 at the lowest X-
ray contour levels, which fits very well with the hook-shaped
morphology of EXE1. Other lanes of absorbing H i gas can be
; (cid:0)33(cid:14))
seen in Figs. 9a and 9b, they run from (`; b) = (207(cid:14)
to (`; b) = (199(cid:14)
; (cid:0)40(cid:14)) to
; (cid:0)48(cid:14)) (both lanes are located close to the (cid:11) = 4h
(`; b) = (204(cid:14)
H(cid:11) filament). The absorption by these H i lanes is visible in
Fig. 9 as a decrease in the X-ray intensity, especially in the
0:25 keV energy band. This suggests that the absorption features
found by BSNGG correspond to kinematically-narrow features
in H i and more clearly establishes the association of the X-ray
enhancement with the Orion-Eridanus Bubble.
; (cid:0)38(cid:14)) and from (`; b) = (197(cid:14)
If the X-ray-emitting gas is located inside an expanding
H i bubble, one would expect the X-rays confined within it to
be absorbed by the approaching, negative-velocity gas. The
receding, positive-velocity gas would only attenuate X-rays
emitted from the general background. For a spherical shell, the
X-ray-emitting gas would be located in front of H i observed at
positive velocities differing little from the expansion velocity.
The anti-correlation between X-ray emission and H i emission
indeed breaks down towards more positive velocities. This is
shown in Fig. 10, where X-ray contours are overlaid on the
gray-scale image of H i emission originating between +20:0 (cid:20)
vLSR (cid:20) +40:0 km s(cid:0)1. The anti-correlation is still significant,
suggesting an elongation of the H i shell along the line of sight.
Contrary to the conclusions of BSNGG, Fig. 9 shows that
the X-ray enhancement is closely associated with the H i cavity
at positive velocities. The H i features associated with the shell
surrounding this cavity can be followed for velocities reaching
up to about +40 km s(cid:0)1, as described in Sect. 3. From the bub-
ble model proposed by BSNGG it follows that most of the H i
emission from the Bubble should be visible at negative veloci-
ties. Our data show the contrary, suggesting that the Bubble is
expanding more symmetrically with respect to the Orion A and
B clouds.
4.3. CO data
The Orion-Eridanus region is too large to have yet been fully
mapped in any molecular-line tracer. Most of the CO data avail-
able in this general area of the sky are confined to the neigh-
borhood of the Orion Molecular Cloud complex. Two large
molecular cloud complexes nearby are the Taurus and Perseus
clouds, but these are located away from the H i shell and may be
ignored in the present context. The OMC was most extensively
covered in 12CO by Maddalena et al. (1986) and in 13CO by
Bally et al. (1987). A comparison between H i and CO in the
region of the Orion A and B clouds was made by Chromey
et al. (1989). At high negative latitudes there is only limited
coverage in the CO data, with only individual clouds studied.
Such data can be found in Dame et al. (1987, down to (cid:0)25(cid:14))
and in Bally et al. (1991, down to (cid:0)30(cid:14)). A number of clouds
detected in the high-latitudeCO survey of Magnani et al. (1985,
henceforth MBM) are located in the Orion-Eridanus area. The
location of these clouds with respect to the H i can be seen in
figures presented by Gir et al. (1994).
Several of the clouds mapped by MBM seem to have clear
counterparts in our H i maps, namely MBM 110, 111, 16, 18,
and 20. Gir et al. (1994) did detailed H i observations towards
10 regions containing 18 MBM clouds. They showed that in
general the H i spectra of the high latitude clouds show sev-
eral components with one component being clearly associated
with the CO cloud. In particular Gir et al. confirm the asso-
ciation of MBM 16, 18 and 20 with H i. A statistical estimate
of the distance to all the clouds in their sample was given
by MBM as about 100 pc. Franco (1988) found a distance
to MBM 18, specifically, of about 130 pc, based on uvby(cid:12)
photometry. Magnani & De Vries (1986) found upper lim-
its to the distances to MBM 18 and MBM 20 of 175 pc and
125 pc, respectively, based on star counts. The two clouds
MBM 110 and MBM 111 are located just below Orion A at
(`; b) = (208(cid:14)
; (cid:0)23(cid:14)) and (209(cid:14),(cid:0)20(cid:14)), respectively. Their ve-
locities are about +10 km s(cid:0)1; a corresponding H i enhance-
ment is seen in Map D. The cloud MBM 16 is located at
(`; b) = (172(cid:14)
; (cid:0)38(cid:14)), just at the edge of the H i shell visi-
ble in Map C. Its velocity is +6:9 km s(cid:0)1; an H i counterpart
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
5
; (cid:0)36(cid:14)) and (`; b) = (211(cid:14)
is visible as an enhancement in Map C. The clouds MBM 18
and MBM 20 are located inside the H i shell seen in Map C at
(`; b) = (189(cid:14)
; (cid:0)37(cid:14)) respectively.
Their velocities are +9:9 km s(cid:0)1 and +0:3 km s(cid:0)1, respectively;
H i counterparts are evident in Map C. (LSR velocities for
MBM 16, 18 and 20 are from Gir et al. 1994.) Note that at the po-
sition of MBM 18 there is also enhanced H i emission in Map B.
This indicates that H i emission seen at negative velocities in the
direction of the Bubble may be due to foreground material. The
cloud MBM 18 is associated with LDN 1569 and located close
to the (cid:11) = 4h filament that appears to absorb the low energy X-
rays; MBM 20 is associated with LDN 1642, in which a number
of pre-main sequence objects is located (Sandell et al. 1987).
Liljestrom & Mattila (1988) mapped LDN 1642 in H i and found
three velocity components at (cid:0)11; +0:5, and +15 km s(cid:0)1, plus
a broad component extending from (cid:0)40 to +50 km s(cid:0)1. This
implies that the foreground material might be present over a
large velocity interval in H i.
In their model BSNGG placed the clouds MBM 18 and 20
between the Local Bubble and the Orion-Eridanus Bubble and
they suggested that these two clouds were formed as a conse-
quence of the interaction of the Local Bubble and the Orion-
Eridanus Bubble. This is consistent with the finding by Gir et al.
(1994) that the H i gas associated with high latitude CO clouds
is mostly in filamentary or loop-like structures. Furthermore
Gir et al. suggest that the molecular clouds are formed in situ,
possibly in the shock compressed gas in shells blown by stellar
winds or supernovae.
There are two clouds visible in H i which have a cometary
morphology. These are located at (`; b) = (203(cid:14)
; (cid:0)32(cid:14)) and at
(206(cid:14),(cid:0)26(cid:14)), and can be seen in both the unsmoothed and fil-
tered versions of the H i data shown in Map B. These structures
are associated with the Lynds Bright Nebula (LBN) 917 and
LBN 959, respectively, which were mapped in CO by Bally
et al. (1991). The velocities are consistent with the H i counter-
part velocities in Map B. From the cometary structure of these
clouds, Bally et al. (1991) infer that they are actually located
inside the Orion-Eridanus Bubble.
Thus we conclude that a substantial fraction of the H i emis-
sion found along the lines of sight toward the central cavity of
the shell seen in Maps B and C is due to foreground material or
to clouds located inside the Bubble.
4.4. IRAS data
; 0(cid:14)) to (`; b) (cid:25) (195(cid:14)
Figure 11a shows a map of the 100 (cid:22)m emission from dust in
the Orion-Eridanus area. It is a mosaic of plates from the IRAS
Sky Survey Atlas (Wheelock et al. 1994). (The sharp border
seen running from (`; b) (cid:25) (220(cid:14)
; (cid:0)30(cid:14))
is due to a slight mismatch in the plate overlap.) A number of
well known clouds, such as Orion A and B, and the (cid:21)-Orionis
ring, can easily be identified in this map. The knot of emission
near (`; b) = (206(cid:14)
; (cid:0)2(cid:14)) is contributed by the Rosette Neb-
ula; although the Rosette locally contaminates our H i emission
maps over the range +10 to +20 km s(cid:0)1 (see Kuchar & Bania
1993), its distance of about 1600 pc places it well beyond the
region of concern in this paper. There is also enhanced emission
with respect to the surroundings at the positions of the clouds
MBM 16, 18, and 20, as well as near LBN 917 and 959. The out-
lines of the H i shell visible in Map C are also seen in the 100 (cid:22)m
map, as well as a lane running from (`; b) = (185(cid:14)
; (cid:0)33(cid:14)) to
(195(cid:14)
; (cid:0)45(cid:14)), which corresponds to the the H(cid:11) filament men-
tioned in section 4.1.
Figure 11b shows a map of the total integrated H i emis-
sion in the Orion-Eridanus area. Note the tight gas-to-dust cor-
relation revealed by the 100 (cid:22)m dust image compared with
the H i data integrated over the entire velocity range. (The cir-
rus/N(H i) correlation has been observed to be quite tight locally,
and throughout the inner Galaxy generally; the correlation does
seem to break down in the outer reaches of the Milky Way,
where H i emission remains intense but the 100 (cid:22)m emission
becomes weak; see Burton 1992.) We note that the cirrus/N(H i)
correlation which appears so tight in Fig. 11 can be kinemati-
cally unravelled, by comparing the dust emission with the H i
contributed from narrow velocity slices of the sort shown in
Fig. 1. Such unravelling has shown that cirrus features can
sometimes be identified with H i emitting at quite anomalous
velocities (see Deul & Burton 1990, and Burton et al. 1991).
Establishing the velocity characteristic of various dust features
in Fig. 11a confirms their association with the Orion-Eridanus
region.
Figure 12 shows the correlation revealed by the H i gas
column densities and the 100 (cid:22)m dust intensities for different
sections of the Orion-Eridanus area. The region that approxi-
mately encloses the H i shell can be delineated by the rectangle
stretching from 170(cid:14)
(cid:20) b (cid:20) (cid:0)30(cid:14).
Figures 12a and 12b show that the gas-to-dust correlation is not
strong outside the region but that it is particularly tight within
it. The latitude strips up to b = (cid:0)30(cid:14) (Figs. 12c and 12d) show
a similar behavior.
(cid:20) ` (cid:20) 230(cid:14) and from (cid:0)55(cid:14)
In studies confined near the Galactic equator, the correlation
between N(H i) and I100(cid:22)m in some cases yields a straight line
where N(H i) gradually flattens off towards higher I100(cid:22)m levels
(see Deul 1988; Burton 1992). This flattening of the correlation
can be explained as due to significant H i optical depths or to
the variation of the interstellar radiation field. In Fig. 12e, there
is such a pronounced flattening which we attribute to increasing
optical depth of the H i. Figures 12a and 12f explicitly include
regions lying quite close to the Galactic equator. Thus, although
for latitudes closer to the equator than b = (cid:0)15(cid:14) the intense H i
emission may not be optically thin, Fig. 12a does show that
the flattening originates from areas outside the H i shell. This
implies that the H i emission from the Orion-Eridanus filaments
and the cavity is optically thin, justifying our determination
of the mass of the shell by directly converting the observed H i
emission into column densities. The gas column depths at a high
level are evidently contributed by gas lying along long lines of
sight, transsecting the Galaxy at large, and therefore no doubt
refer to gas not relevant to our discussion of the Orion-Eridanus
Bubble.
6
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
5. Mass, energetics, and origin of the Orion-Eridanus
Bubble
5.1. Mass of the Bubble
In order to determine the total mass of the H i in the shell sur-
rounding the Orion-Eridanus Bubble we proceeded as follows.
We isolated all the filaments and loops that we identify as part
of the shell (see section 3) by marking those positions on the
sky that coincide with the identified features. The filaments and
loops that we identify as part of the expanding shell are shown
in Fig. 8. In all positions coincident with the filaments the col-
umn densities were determined from the observed H i-emission
Tb(cid:1)v by 1:8224 (cid:2) 1018 (K km=s)(cid:0)1cm(cid:0)2.
by multiplying P
The observed column density at each position on the sky can
then be converted into an H i mass.
Prior to the determination of the column density we sub-
tracted a smooth background from each map to take into account
the H i emission from regions along the line of sight that are
not associated with the shell. The background was determined
by fitting bivariate polynomials to the H i maps or by simply
determining a constant background level. The total mass one
finds differs depending on the background subtraction (the de-
gree of the fitting polynomial). By determining the mass using
the different background subtractions one gets a handle on the
uncertainty in the mass estimate. The mass we determined is
2:3(cid:6)0:7(cid:2)105
2
380 M(cid:12), where d380 is the distance to the Orion-
Eridanus Bubble in units of 380 pc. Our mass estimate includes
a factor of 1:4 to account for the total-mass to H i-mass ratio.
d
Further causes of uncertainty in the mass estimate of the
shell are: possible misidentification of some features in the H i
maps, errors made in isolating the features, and the distance to
the features (which were now all placed at 380 pc).
d
d
The total H i-mass estimates of the Bubble made by other
2
authors are 4:7(cid:2)105
2
380 M(cid:12) (Heiles 1976), 2:6(cid:2)104
380 M(cid:12)
(BSNGG), and 4:3(cid:2)103
2
380 M(cid:12) (BSNGG, using the negative-
velocity cap). We note that the masses found by BSNGG are
considerably smaller, due to the smaller size they infer for the
shell surrounding EXE1.
d
5.2. Energy and origin of the Bubble
Following the discussion in Sect. 5.2 of Paper I, we investigate
whether the Bubble could have been formed by the stellar winds
and supernovae from Orion OB1. In the simplest approxima-
tion we assume the shell to be spherical and in that case the
observed velocity extent of the shell corresponds to its expan-
sion velocity. From our estimate of its mass, 2:3 (cid:2)105 M(cid:12), and
its expansion velocity, 40 km s(cid:0)1, we can calculate the kinetic
energy of the expanding H i shell. If we assume that the bubble
is expanding uniformly at this velocity then the kinetic energy is
3:7 (cid:2) 1051 ergs. We followed the model of wind-driven bubbles
given by Castor et al. (1975) and Weaver et al. (1977), assuming
a spherical shell embedded in a homogeneous ambient medium.
The relevant equations for the shell radius, RS, and expansion
velocity, VS, are those given by Shull (1993) (see Paper I).
This model predicts that some 20% of the total power re-
leased into the ISM by an OB association is converted into
kinetic energy of the shell surrounding the wind-blown bub-
ble. This implies that the Orion-Eridanus Bubble requires a
total energy output over the lifetime of Orion OB1 of about
1:9 (cid:2) 1052 ergs. According to Paper I the total energy output
from Orion OB1 is approximately 1052 ergs. This number in-
cludes contributions from both supernovae and stellar winds.
We note that in modelling the evolution of the shell the mechan-
ical luminosity from the association is considered constant in
time. Thus the supernova contribution is averaged over the life-
time of the association and no distinction between a wind-driven
or supernova-driven bubble is made. However, both contribu-
tions are important, with the stellar winds contributing 25% of
the total energy output in the case of Orion OB1 (see Paper I).
The total energy output from Orion OB1 is less than the energy
required here by a factor of 2. The discrepancy may be due to
the uncertainties in the estimated mass of the shell. It seems
at least plausible from the energy requirements that the Orion-
Eridanus Bubble has been blown into the ISM by the stellar
winds and supernovae in Orion OB1.
From the estimated mass of the Bubble and its observed
size we derive the initial ambient density of the ISM to be about
0:9 cm(cid:0)3. This is a large value at such high jz j. The H i volume-
density value of 0:5 cm(cid:0)3 found generally characteristic close
to the Galactic equator does, however, represent averaging over
Galaxy-scale volumes of space. Such a general value may not
be valid for the Orion-Eridanus region which is known to be one
of intense star formation and therefore will be a site of generally
high densities. We note that the mass estimate is uncertain and,
furthermore, that the assumption of a spherical geometry for
the Bubble may have led to an underestimate of its volume.
(If an apparently spherical shell is elongated along the line of
sight its volume will be larger than in the spherical case.) As
discussed in Paper I, we can calculate RS and VS using only the
energy output from Subgroup 1a of Orion OB1. If we assume
that the age of the Bubble equals the age of this subgroup
(11:4 Myr, see Paper I), the energy output from Subgroup 1a
is about 2 (cid:2) 1037 ergs s(cid:0)1. This yields a value for RS which is
larger (210 pc) and a value for VS which is lower (11 km s(cid:0)1)
than the values which we found above. For an initial ambient
density of half the value of 0:9 cm(cid:0)3, these numbers are 240 pc
and 13 km s(cid:0)1, respectively. If we demand that the age be
consistent with our observed values of RS and VS, it follows
that the age of the Bubble is (cid:24) 1:8 Myr and the corresponding
mechanical luminosity is (cid:24) 4 (cid:2) 1038 ergs s(cid:0)1. This value for
the age of the shell is significantly different from the age of the
oldest subgroup in Orion OB1, and the required luminosity is
too high.
We note, however, that Orion OB1 is not located in a ho-
mogeneous ambient medium and that the association is located
off-center inside the bubble (see Fig. 7). The first step in re-
fining the model for the bubble is to take into account the fact
that on average the Galactic H i layer is stratified perpendic-
ular to the plane of the Galaxy. In the simplest approach the
Galactic H i layer is approximately plane parallel. We would
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
7
like to stress here that this stratification holds only on aver-
age. In localized regions (such as the Orion-Eridanus region) in
the Galaxy one can find large density enhancements or under-
dense regions. Density variations that existed before the stellar
winds and supernovae occurred will affect the expansion of the
bubble. Comparison of a model which only takes the plane par-
allel density stratification into account to the observations will
provide insight into these initial density inhomogeneities.
No simple analytical results have been derived for H i bub-
bles expanding in plane parallel stratified media, but various
numerical simulations have been performed (see Tomisaka &
Ikeuchi 1986; Mac Low & McCray 1988; Tenorio-Tagle & Bo-
denheimer 1988). For the case of adiabatic bubbles expanding
into spherically symmetric media containing a finite amount of
mass (implying density stratification), a semi-analytical treat-
ment has been given by Koo & McKee (1990). Their model can
be generalized to the plane parallel case (see below). The effects
of the plane-parallel geometry of the Galactic H i are that bub-
bles will grow much larger in the direction perpendicular to the
plane and might experience blow out (see Mac Low & McCray
1988). If the association that causes the bubble is located above
or below the Galactic plane then the bubble will grow larger in
the direction away from the plane. This explains the off center
location of Orion OB1 with respect to the bubble. Note that in
a plane parallel medium the expansion of the shell will not take
place at a constant velocity over its surface. This means that the
observed extent in velocity space no longer corresponds to the
expansion velocity but rather indicates the lower limit on the
actual expansion velocities in the shell surrounding the bubble.
In finite (stratified) media the shell surrounding the bub-
ble will eventually accelerate and possibly blow out from the
Galactic disk. Throughout the evolution of the bubble the ob-
served expansion velocities will be higher than expected on the
basis of the theory for a uniform medium. The models by Koo &
McKee (1990) allow the calculation of the evolution of the shell
of swept up matter surrounding the expanding bubble, given the
central density and the mechanical energy input. The calcula-
tions for the spherically-symmetric case can be generalized to
plane-parallel media, by assuming that all mass elements in the
shell move on radial tracks (sector approximation) away from
the source of energy. The equations for the one-dimensional
case are assumed to remain valid. This implies that along the
z-axis the results are the same as for the one-dimensional case,
and along the direction perpendicular to the z-axis the solution
will be the same as for a uniform medium. The approximation
used by Koo & McKee underestimates the sizes of the bubbles,
but is accurate to within 10% prior to the acceleration of the
shell, as compared to numerical calculations. Once the shell has
accelerated, the approximation is accurate to within 20%.
The Orion-Eridanus Bubble is located close to the Sun and
its size is comparable to its distance. This means that projection
effects are very important when one is observing the Bubble.
Thus, when modelling the Bubble one should construct a three-
dimensional model for the shell surrounding the Bubble and
project the shell onto the sky in order to compare the model
to the observations. We do this by calculating models with
the Koo & McKee formalism for the two-dimensional case and
rotating the resulting model around the axis perpendicular to the
Galactic plane. The expansion center of the model (the Orion
OB1 association) is placed at 380 pc from the Sun and 124 pc
below the Galactic plane at a Galactic longitude of 205(cid:14). The
models are compared to the observations under the requirement
that the high latitude projection of the model matches the shell
structures at b (cid:25) (cid:0)50(cid:14).
The Galactic H i layer has been described as consisting of
two Gaussian components, with FWHM of 212 and 530 pc, and
a broad exponential component with a scale height of 403 pc.
The densities in the Galactic plane of these components are
0:395, 0:107, and 0:064 cm(cid:0)3, respectively (Dickey & Lock-
man 1990). Calculation of the expected z-profile of the Galactic
H i layer in a realistic potential was done by Malhotra (1994).
She shows that the result can approximately be described by
a Gaussian z-distribution with a scale height of 144 pc and a
mid-plane density of 0:86 cm(cid:0)3. Both density distributions are
used in the model calculations. The results are shown in Fig.
13. In both cases we show map C overlaid by the model. The
model consists of a grid of points on the surface of the shell and
is projected into an (`; b; v) data cube with the same spatial and
kinematical resolution as the H i data cube. The overlaid model
points are all those points on the surface of the shell that can
be observed in the velocity interval corresponding to map C.
Note that we do not model the thickness of the swept up shell.
The irregular structure in the model data is due to the finite
resolution with which the surface of the shell is represented.
When we calculate the shell evolution for the Dickey &
Lockman density distribution the best match to the b (cid:25) (cid:0)50(cid:14)
filaments is for an age of 5:3 Myr. The velocity range spanned
in the projected model is (cid:0)19 to +24 km s(cid:0)1 and the actual
maximum velocity is 29 km s(cid:0)1 (note that due to projection
effects the model shell is not symmetrical in velocity space).
The bottom of the shell is located at 330 pc below the Galactic
plane according to this model. For the model in which the
Gaussian density distribution is used the best match is at an
age of 5:8 Myr. The velocity range spanned in the projected
model is (cid:0)17 to +23 km s(cid:0)1 and the actual maximum velocity
is 29 km s(cid:0)1. The bottom of the shell is located at 335 pc below
the Galactic plane.
In both cases the model shell is much too large in the direc-
tion of the Galactic plane, as well as towards large longitudes.
This indicates that indeed the assumed density stratifications
are not appropriate for the Orion-Eridanus area. The Orion
OB1 association is located close to the Orion molecular clouds
and they probably constrain the expansion of the shell. The
expansion also seems to be halted by density enhancements lo-
cated between the Orion and Taurus molecular clouds (around
(`; b) (cid:25) (180(cid:14)
; (cid:0)20(cid:14))). And indeed there are large H i column
densities in that region. The expansion seems to be halted also
by density enhancements along the bright ridge that runs from
(`; b) = (170(cid:14)
; (cid:0)24(cid:14)) in Map C. The
observed shell is, however, larger than the models if one looks
at the lower longitude side. This may indicate the presence of
; (cid:0)40(cid:14)) to (`; b) = (180(cid:14)
8
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
an underdense region around (`; b) (cid:25) (185(cid:14)
the shell expands more easily.
; (cid:0)40(cid:14)), into which
One could evolve the models further in time to have them
; (cid:0)50(cid:14)), but that
match the loop structure at (`; b) (cid:25) (190(cid:14)
would require a confinement of the expansion in the direc-
tion (230(cid:14)
; (cid:0)50(cid:14)). There are no obvious density enhancements
there.
When looking at higher velocities the loops associated with
the shell should decrease in size and eventually form a closed
cap. But if we compare our model calculations to the obser-
vations at higher velocities we see that the decrease in size is
much more rapid for the model than for the observed shell.
This is already indicated in the channel maps of figure 1, where
the elliptic structure at positive velocity shows no significant
change in size. This discrepancy suggest that the observed shell
is elongated along the line of sight. This was already suggested
by BSNGG (see also section 4.2).
The dynamical age of the Bubble is consistently lower than
the age of the oldest subgroup in Orion OB1. This may be
explained by considering that a young subgroup first had to dis-
rupt its parental molecular cloud, through the combined action
of ionizing radiation and stellar winds, before it could blow a
bubble into the ambient medium. In that case, not all of the
energy from Subgroup 1a would have gone into the Bubble,
but the energy from Subgroups 1b and 1c will have contributed
as well. Furthermore, we have assumed a constant mechanical
luminosity for the Orion OB1 association. In reality the energy
output is not constant and depends on the star formation his-
tory of the association. As shown by Shull & Saken (1994),
in the case of coeval or continuous star formation a peak in
the energy output will occur as massive stars evolve off the
main sequence. The peak in the output can be 3 -- 10 times the
later energy deposition rate due to supernovae and Wolf-Rayet
stars. This peak in the energy output rate will result in faster
shell growth and higher expansion velocities than in models
with constant luminosity. Furthermore, they show that the ages
of bubbles are underestimated by constant luminosity models.
The underestimate is especially severe in the case of non-coeval
star formation, an example of which is the formation of sub-
groups in an association. These findings are consistent with the
fact that the age of the shell according to our model is lower
than the age of the association and that the observed expansion
velocities are higher than the model velocities.
Another way in which mechanical energy from Orion OB1
may be deposited into the ISM is through runaway OB stars.
There are three well known runaway stars associated with Orion
OB1: (cid:22) Col (O9:5V), AE Aur (O9:5V) and 53 Ari (B2V)
(Blaauw 1961, 1993). The space velocities of these stars rela-
tive to Orion OB1 are estimated to be 117, 140 and 40 km s(cid:0)1,
respectively. The time elapsed since their ejection from the cen-
ter of Orion OB1 would be 2:3, 2:4 and 7:3 Myr. This implies
that early type runaways can travel distances comparable to the
size of the Orion-Eridanus Bubble in short times. It is possi-
ble that there were more runaways in the past and that they
have already exploded as supernovae. One supernova explo-
sion (with an explosion energy of 1051 ergs, and located in the
same plane parallel media as used above) in the center of the
Orion-Eridanus Bubble (at (`; b) (cid:25) (195(cid:14)
; (cid:0)35(cid:14))) could cause
a shell with the same size and velocity extent as the observed
shell. However the swept up mass would be too small by an
order of magnitude compared to the observed shell mass. Let-
ting the supernova explode in a uniform medium of density
0:9 cm(cid:0)3 (see above) would obviously result in the right swept
up mass, but the expansion velocity of the shell would be too
low. However, the occurrence of a supernova outside the shell
due to a runaway could create a low density environment into
which the shell expands more easily and at higher velocities.
Alternatively, the runaway could explode far from the associ-
ation in an already existent bubble and will then provide extra
energy for the shell expansion from a position close to the edge
of the shell. This could be the cause of observed high veloc-
ities, where parts of the shell have been re-accelerated by the
supernova explosion.
We conclude that the size and expansion velocity of the
Orion-Eridanus Bubble can be explained as the result of stellar
winds and supernovae from Orion OB1.
6. Conclusions
We presented maps of neutral hydrogen emission from the area
surrounding the Orion-Eridanus Bubble, and showed that the
observed velocity extent of the Bubble is about (cid:6)40 km s(cid:0)1.
The enhanced X-ray emission in this area of the sky anti-
correlates in a detailed way with kinematically-narrow features
in H i, clearly establishing the association of the X-ray enhance-
ment with the Bubble. Based on published CO data we conclude
that most of the emission observed in the direction where a cav-
ity is expected in Maps B and C originates from foreground
clouds. Evidently, the observed X-ray emission is associated
with the inside of a single, coherent H i shell. From comparison
of the IRAS 100 (cid:22)m dust data with the H i data we concluded
that the H i shell is optically thin, which allowed us to deter-
mine the mass of the shell. The mass of the shell surrounding
the Bubble was found to be 2:3 (cid:6) 0:7 (cid:2) 105 M(cid:12). Assuming
spherical symmetry, the mass and the expansion velocity of
the shell correspond to a kinetic energy to be 3:7 (cid:2) 1051 ergs.
According to the standard model the energy output from Orion
OB1 is just large enough to account for the kinetic energy of
the Bubble. Using models for wind-blown bubbles which in-
corporate the density stratification of the Galactic H i layer, we
showed that the Orion-Eridanus feature can be explained as a
wind-blown bubble. However, density inhomogeneities in the
ambient medium cause large discrepancies between our model
and the observed shell. The energy source for this bubble is the
mechanical luminosity from stellar winds and from supernovae
in Orion OB1.
Acknowledgements. We thank D.N. Burrows for kindly pro-
viding us the BSNGG X-ray data used in this paper. We are
grateful for discussions with Tim de Zeeuw, Eugene de Geus
and Adriaan Blaauw. We thank the referee for comments that
helped improve the final version of this paper. The Infrared
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
9
Heiles, C., 1984, ApJS 55, 585
Heiles, C., Habing, H.J., 1974, A&AS 14, 1
Heiles, C., Jenkins, E.B., 1976, A&A 46, 333
Kalberla, P.M.W., 1978, Die Korrektur der Linienprofile der
21-cm Emission bezuglich der Streustrahlung des Anten-
nendiagrams, PhD. thesis (in German; English translation
by R.J. Cohen), University of Bonn
Kalberla, P.M.W., Mebold, U., Reich, W., 1980, A&A 82, 275
Koo, B.-C., McKee, C.F., 1990, ApJ 354, 513
Kuchar, T.A., Bania, T.M., 1993, ApJ 414, 664
Kutner, M.L., Tucker, K.D., Chin, G., Thaddeus, P., 1977, ApJ
215, 521
Liljestrom, T., Mattila, K., 1988, A&A 196, 243
Mac Low, M.-M., McCray, R., 1988, ApJ 324, 776
Maddalena, R.J., Morris, M., Moscowitz, J., Thaddeus, P., 1986,
ApJ 303, 375
Magnani, L., de Vries, C.P., 1986, A&A 168, 271
Magnani, L., Blitz, L., Mundy, L., 1985, ApJ 295, 402 (MBM)
Malhotra, S., 1994, ApJ submitted
Malin, D.F., 1979, Mercury 8, 89
Reynolds, R.J., Ogden, P.M., 1979, ApJ 229, 942 (RO)
Reynolds, R.J., Roesler, F.L., Scherb, F., 1974, ApJ 192, L53
Sandell, G., Reipurth, B., Gahm, G., 1987, A&A 181, 283
Shull, J.M., 1993, in Massive Stars: Their Lives in the Interstel-
lar Medium, eds. J.P. Cassinelli and E.B. Churchwell, ASP
conference series Vol. 35, p. 327
Shull, J.M., Saken, J.M., 1994, ApJ submitted
Sivan, J.P., 1974, A&AS 16, 163
Stark, A.A., Gammie, C.F., Wilson, R.W., Bally, J., Linke, R.A.,
Heiles, C., Hurwitz, M., 1992, ApJS 79, 77
Tenorio-Tagle, G., Bodenheimer, P., 1988, ARA&A 26, 145
Tomisaka, K., Ikeuchi, S., 1986, PASJ 38, 697
Weaver, H.F., Williams, D.R.W., 1973, A&AS 8, 1
Weaver, R., McCray, R., Castor, J., Shapiro, P., Moore, R., 1977,
ApJ 218, 377 (err 220, 742)
Wheelock, S.L., Gautier, T.N., Chillemi, J., et al., 1994, IRAS
Sky Survey Atlas Explanatory Supplement, Jet Propulsion
Laboratory, Pasadena
Astronomical Satellite (IRAS) was a joint project of NASA
(U.S.), NIVR (The Netherlands), and SERC (U.K.). This re-
search was supported in part by the Netherlands Foundation for
Research in Astronomy (NFRA) with financial support from
the Netherlands Organization for Scientific Research (NWO).
References
Bally, J., Langer, W.D., Stark, A.A., Wilson, R.W., 1987, ApJ
312, L45
Bally, J., Langer W.D., Wilson, R.W., Stark, A.A., Pound, M.W.,
1991, in Fragmentation of Molecular Clouds and Star For-
mation, IAU Symp. 147, eds. E. Falgarone, F. Boulanger &
G. Duvert, Kluwer, Dordrecht, p. 11
Blaauw, A., 1964, ARA&A 2, 213
Blaauw, A., 1961, Bull. Astr. Inst. Netherlands 15, 265
Blaauw, A., 1993, in Massive Stars: Their Lives in The In-
terstellar Medium, eds. J.P. Cassinelli & E.B. Churchwell,
ASP conference series, Vol. 35, p. 207
Brown, A.G.A., de Geus, E.J., de Zeeuw, P.T., 1994, A&A 289,
101 (Paper I)
Burrows, D.N., Singh, K.P., Nousek, J.A., Garmire, G.P., Good,
J., 1993, ApJ 406, 97 (BSNGG)
Burton, W.B., 1992, in The Galactic Interstellar Medium, W.B.
Burton, B.G. Elmegreen, R. Genzel; edited by D. Pfenniger
& P. Bartholdi (Saas-Fee Advanced Course 21; Lecture
Notes 1991), Springer-Verlag, Heidelberg, p. 1
Burton, W.B., Bania, T.M., Hartmann, D., Yuang, T., 1991, in
Evolution of Interstellar Matter and Dynamics of Galaxies,
eds. J. Palous, W.B. Burton & P.O. Lindblad, Cambridge
University Press, p. 25
Castor, J., McCray, R., Weaver, R., 1975, ApJ 200, L107
Chromey, F.R., Elmegreen, B.G., Elmegreen, D.M., 1989, AJ
98, 2203
Cowie, L.L., Songaila, A., York, D.G., 1979, ApJ 230, 469
Dame, T.M., Ungerechts, H., Cohen, R.S., de Geus, E.J., Gre-
nier, I.A., May, J., Murphy, D.C., Nyman, L.- A., Thaddeus,
P., 1987, ApJ 322, 706
Deul, E.R., 1988, Interstellar Dust and Gas in the Milky Way
and M33, PhD. thesis, University of Leiden
Deul, E.R., Burton, W.B., 1990, A&A 230, 153
Dickey, J.M., Lockman, F.J., 1990, ARA&A 28, 215
Franco, G.A.P., 1988, A&A 202, 173
Gir, B.-Y., Blitz, L., Magnani, L., 1994, ApJ 434, 162
Goudis, C., 1982, The Orion Complex: A Case Study of Inter-
stellar Matter, Astrophys. Space Sci. Libr., Vol. 90, Reidel,
Dordrecht
Green, D.A., 1991, MNRAS 253, 350
Green, D.A., Padman, R., 1993, MNRAS 263, 535
Hartmann, D., 1994, The Leiden/Dwingeloo Survey of Galactic
Neutral Hydrogen, PhD. thesis, University of Leiden
Hartmann, D., Burton, W.B., 1995, An Atlas of the Galactic
Neutral Hydrogen Emission, Cambridge University Press,
in preparation
Heiles, C., 1976, ApJ 208, L137
10
A.G.A. Brown, D. Hartmann and W.B. Burton: The Orion-Eridanus Bubble
Fig. 1. Caption is included beneath Fig. 1
[Map A] Integrated H i emission in the velocity interval
Fig. 2.
(cid:0)40 km s(cid:0)1
(cid:20) vLSR (cid:20) (cid:0)30 km s(cid:0)1. Both a logarithmically scaled
image (left) and an unsharp-masked representation (right) are shown.
Filaments located between b = (cid:0)50(cid:14) and b = (cid:0)30(cid:14) correspond to the
approaching side of the expanding H i bubble
[Map B] Integrated H i emission in the velocity interval
Fig. 3.
(cid:0)28 km s(cid:0)1
(cid:20) vLSR (cid:20) (cid:0)3 km s(cid:0)1. The logarithmically scaled im-
age is shown on the left and the unsharp-masked representation on the
right. The major portion of an H i shell is visible between b = (cid:0)50(cid:14)
and b = (cid:0)20(cid:14). The emission inside this structure may well be largely
contributed by foreground material
[Map C] Integrated H i emission in the velocity interval
Fig. 4.
(cid:0)1 km s(cid:0)1
(cid:20) vLSR (cid:20) +8 km s(cid:0)1. The logarithmically scaled image is
shown on the left and the unsharp-masked representation on the right.
The largest coherent structure in Orion-Eridanus can be seen in these
maps. The unsharp-masked image shows the loop between ` = 170(cid:14)
and ` = 230(cid:14) and b = (cid:0)50(cid:14) and b = (cid:0)17(cid:14). The Orion clouds A and
B are located near (`; b) = (208(cid:14)
; (cid:0)16(cid:14)). The (cid:21)-Orionis ring is visible
around (`; b) = (195(cid:14)
; (cid:0)12(cid:14)) (see also Fig. 11)
[Map D] Integrated H i emission in the velocity interval
Fig. 5.
+9 km s(cid:0)1
(cid:20) vLSR (cid:20) +14 km s(cid:0)1. The logarithmically scaled im-
age is shown on the left and the unsharp-masked representation on
the right. The filament running between ` = 170(cid:14) and ` = 220(cid:14) and
b = (cid:0)53(cid:14) and b = (cid:0)20(cid:14) has a large velocity extent. It can also be
associated with the loop structure in Map C
Fig. 6.
[Map E] Integrated H i emission in the velocity interval
+15 km s(cid:0)1
(cid:20) vLSR (cid:20) +40 km s(cid:0)1. The logarithmically scaled im-
age is shown on the left and the unsharp-masked representation on
the right. The elliptical structure between ` = 170(cid:14) and ` = 215(cid:14) and
b = (cid:0)53(cid:14) and b = (cid:0)30(cid:14) can be traced in Fig. 1 up to vLSR = +40 km s(cid:0)1;
it represents the receding side of the Orion-Eridanus H i bubble
Fig. 7. The position of the Orion OB1 association with respect
to the H i shell. The grey scale image is a logarithmically scaled
representation of integrated H i emission in the velocity interval
(cid:0)1 km s(cid:0)1
(cid:20) vLSR (cid:20) +8 km s(cid:0)1. The contours outline the 100 (cid:22)m
(IRAS) emmission from the Orion A and B molecular clouds (the ring
around (`; b) = (195(cid:14)
; (cid:0)12(cid:14)) is the (cid:21)-Orionis ring). The dots show the
brightest stars in the Orion constellation. The circles show the posi-
tions of the three main subgroups of Orion OB1. From right to left are
shown 1a, 1b and 1c
Fig. 8. Bitmaps, corresponding to Figs. 2 -- 5, that show the features
that are identified as part of the expanding H i shell in (a) map A, (b)
map B, (c) Map C, (d) map D
Fig. 8e. Bitmap, corresponding to Fig. 6, that shows the features that
are identified as part of the expanding H i shell in map E.
Fig. 9. Anti-correlation of H i emission (gray-scale image) and X-ray
(contours) emission. The H i data are logarithmically scaled, and the
contour values for the X-ray data are in counts s(cid:0)1. (a) H i emission
between (cid:0)1 (cid:20) vLSR (cid:20) +4 km s(cid:0)1 and 0:25 keV (L1 band) X-ray
emission. (b) H i as in (a) and 0:6 keV (M1 band) X-ray emission. (c)
H i emission between +13:5 (cid:20) vLSR (cid:20) +18:5 km s(cid:0)1 and 0:25 keV
X-ray emission. (d) H i as in (c) and 0:6 keV X-ray emission
Fig. 10. Anti-correlation between X-ray emission (contours,
in
counts s(cid:0)1) and H i emission (gray-scale image, logarithmically scaled)
for the velocity interval +20 (cid:20) vLSR (cid:20) +40 km s(cid:0)1. Both the 0:25 keV
(L1 band) emission (a) and the 0:6 keV (M1 band) emission (b) display
a prominent anti-correlation with the positive-velocity H i gas from the
lower-b region of the H i cavity; the H i gas lying closer to the Galactic
equator may well be ambient foreground material, not associated with
the Orion-Eridanus Bubble
Fig. 11. Maps of the extended Orion-Eridanus region showing in (a)
the 100 (cid:22)m IRAS dust emission and in (b) the total integrated H i
emission. The generally-tight correlation is considered in the text in
specific velocity- and spatial regions
Fig. 12. Scatter diagrams of N(H i) vs. I100(cid:22)m for different regions in
the Orion-Eridanus area, shown separately in (a) for the region outside
the approximately rectangular boundary of the shell, and in (b) for
the region inside the shell. Panels (c) -- (f) show in 15(cid:14) wide latitude
strips that the linear correlation between N(H i) and 100 (cid:22)m emission
is strong at b < (cid:0)30(cid:14), but breaks down nearer the Galactic plane
Fig. 13. Map C overlaid with the model for the expanding H i shell
for: (a) an H i layer with a Gaussian density distribution (scale height
144 pc and mid-plane density 0:86 cm(cid:0)3) and (b) an H i layer with a
density distribution as given by Dickey & Lockman (1990). In both
cases we show all the points on the surface of the shell model that
can be observed in the velocity interval corresponding to map C. The
irregular structure of the models is due to finite resolution with with
the shell surface is represented
This article was processed by the author using Springer-Verlag TEX
A&A macro package 1992.
|
astro-ph/9808044 | 1 | 9808 | 1998-08-06T06:02:56 | Mapping the CMB II: the second flight of the QMAP experiment | [
"astro-ph"
] | We report the results from the second flight of QMAP, an experiment to map the cosmic microwave background near the North Celestial Pole. We present maps of the sky at 31 and 42 GHz as well as a measurement of the angular power spectrum covering the l-range 40-200. Anisotropy is detected at about 20 sigma and is in agreement with previous results at these angular scales. We also report details of the data reduction and analysis techniques which were used for both flights of QMAP. | astro-ph | astro-ph | Submitted to Astrophysical Journal Letters August 5, 1998
Preprint typeset using LATEX style emulateapj v. 04/03/99
8
9
9
1
g
u
A
6
1
v
4
4
0
8
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
MAPPING THE CMB II: THE SECOND FLIGHT OF THE QMAP EXPERIMENT
T. Herbig1,4, A. de Oliveira-Costa1,2, M. J. Devlin3, A. D. Miller1, L. Page1, and
Submitted to Astrophysical Journal Letters August 5, 1998
M. Tegmark2,4
ABSTRACT
We report the results from the second flight of QMAP, an experiment to map the cosmic microwave
background near the North Celestial Pole. We present maps of the sky at 31 and 42 GHz as well as a
measurement of the angular power spectrum covering 40 ∼< ℓ ∼< 200. Anisotropy is detected at ∼> 20σ
and is in agreement with previous results at these angular scales. We also report details of the data
reduction and analysis techniques which were used for both flights of QMAP.
Subject headings: cosmic microwave background-cosmology: observations
1.
INTRODUCTION
The QMAP experiment is a balloon-borne telescope de-
signed to map the cosmic microwave background (CMB)
from angular scales of 0.◦7 to several degrees. Its design
is similar to that of the Saskatoon experiment (Wollack et
al. 1997), but the absence of significant atmospheric emis-
sion at balloon altitudes and the interlocking scan pattern
allow the production of a sky map in an area close to the
North Celestial Pole (NCP). QMAP flew twice in 1996.
The telescope has 6 radiometry channels in 3 dual-
polarization beams. Two of these channels are at Ka-band
(31 GHz) and four are at Q-band (42 GHz). The telescope
can be pointed in azimuth, but the beams have a fixed
elevation of about 41◦. The three beams are chopped si-
nusoidally in azimuth at 4.6 Hz over an angle of up to
θc = 20◦ on the sky. In addition, the azimuth of the tele-
scope is "wobbled" sinusoidally with periods of 50–100 s
and amplitudes of approximately half the chop amplitude.
This Letter reports the results from the second flight of
QMAP. It also details the calibration and pointing anal-
ysis common to both flights of the experiment. A com-
panion paper (Devlin et al. 1998, hereafter D98) presents
the results of flight 1 and discusses the design and perfor-
mance of the instrument. Another companion paper (de
Oliveira-Costa et al. 1998; dO98) presents the combined
results from both flights and explains the mapping tech-
nique and the method used to calculate the angular power
spectrum of the CMB.
2. OBSERVATIONS AND DATA
The second flight of
the QMAP experiment was
launched 1996 November 10 at 23:05 UT from Ft. Sumner,
NM, by the National Scientific Balloon Facility (NSBF).
The gondola reached its float altitude of 30 km two hours
later and stayed at float until 12:55 UT. Science and cali-
bration data were taken betweem 01:00 and 12:10 UT, all
during the night. The flight was terminated 350 km east,
near Shamrock, TX.
1Princeton University, Physics Department, Jadwin Hall,
Princeton, NJ 08544; [email protected]
2Institute for Advanced Study, Olden Lane, Princeton, NJ 08540
3University of Pennsylvania, Department of Physics and
Astronomy, David Rittenhouse Lab., Philadelphia, PA 19104
4Hubble Fellow
In contrast to flight 1, this flight was optimized for
a higher signal/noise ratio for each map pixel; this was
achieved with a smaller chop of θc = 5◦ and a wobble of
∆az ≈ 3◦ at a period of 50 s. Because the latitude of the
flight was ≈ 3◦ higher than flight 1, the resulting maps
are closer to the NCP. The sky coverage for both flights
is shown in D98. The target region around the NCP was
observed in 3 segments of 17305 s, 9200 s, and 3974 s du-
ration, interrupted by celestial calibrations. At the chop
frequency of 4.6 Hz and the data rate of 160 samples/chop,
22,433,920 radiometry samples on the NCP were gathered
in each channel. All 6 channels functioned cleanly-no
data had to be excised because of excess noise.
During ascent, the attitude computer froze repeatedly,
which delayed starting the radiometer cooling system and
prevented the radiometer from reaching thermal equilib-
rium. The resulting amplifier gain variations were tracked
and corrected using periodic calibration pulses. Also
during ascent, the data acquisition computer froze for
≈ 50 s but resumed functioning without rebooting. Con-
sequently, the data buffers overflowed, resulting in large
timing offsets between radiometry, chopper, and pointing
samples. This complicated the phase solutions described
in section 3.2. No other timing delays were encountered.
The primary CCD pointing camera had an 8◦ field of
view, while the secondary camera covered 12◦; both guided
with < 10′′ intrinsic pointing noise at 10 Hz. Almost all
observations were made using the narrow-field camera. As
discussed in section 3.1, no stars were observed in the CCD
during the radiometer calibration on Cas A, resulting in
a large azimuth pointing uncertainty. Two ballast drops
during the second NCP segment, tracked well in the point-
ing, had no discernible effect on the data.
Apart from the radiometer, all system components
reached thermal regulation quickly. Exposed passive ele-
ments, such as the chopper and the mirror, cooled to tem-
peratures of ≈ −45 C upon reaching float altitude, and
stayed within 10 K for the remainder of the flight. The
gondola position was recorded every minute by the GPS
receivers in the NSBF flight package; time was recorded
to 2 s absolute accuracy with the on-board clock.
3. ANALYSIS
The analysis procedure is similar for both flights of
QMAP; the key differences are a degraded CCD focus in
1
2
QMAP 2
flight 1 and large phase delays in flight 2.
3.1. Pointing
For both flights, the CCD camera was the primary
pointing device. For each second of data, the az/el posi-
tions of the guide stars and planets were calculated using
the on-bard clock and the GPS position of the telescope.
The camera scale was calibrated by measuring the posi-
tions of several stars visible in the camera field while the
telescope was held at constant azimuth using the mag-
netometer. From laboratory measurements of the chop-
per and camera orientations and our limits on the camera
horizon during flight, we find that the maximum devia-
tion of the beam at the endpoints of the maximum chop
(θc = 20◦) is < 0.◦1 in elevation.
The offset of the CCD center from the radio beams was
determined directly in flight 1, where a guide star was
available while the radio source Cas A was observed. Dur-
ing flight 2, no guide star was found (Cas A was setting
instead of rising), and the radio center (at az = −39◦)
had to be transferred to the CCD center (at the NCP)
using the magnetometer. The magnetometer was cal-
ibrated using CCD star positions at various azimuths
(−0.◦8 < az < +138.◦1, which excludes the azimuth of
Cas A) and a global fit during a full spin of the telescope.
From this, we estimate the remaining uncertainty in the
radio positions at 0.◦5 in azimuth, or 0.◦4 on the sky. The
resulting smearing of map pixels is minor: depending on
the channel, the average smearing is 0.◦06, with only 10%
of Ka1/2 and Q1/2 samples smeared by more than 0.◦12
(0.◦19 for Q3/4). No section of the map is smeared more
than 0.◦2 (0.◦3 for Q3/4), small compared to the beam sizes.
The pointing analysis of flight 1 was complicated by
a defocussed CCD camera, which resulted in 5′ pointing
noise at 10 Hz as well as distortions of the CCD camera
field of ≈ 0.◦5 at the edges. The pointing noise could be
smoothed easily, while the field distortions caused the az-
imuth wobble to appear as elevation nodding at the wob-
ble frequency and its first harmonic. From observations
of Cas A we found this nodding to be spurious and sup-
pressed it for subsequent analysis.
Additional pointing effects include elevation drifts at
τ ∼ 300 s of ∆el ∼ 0.◦1 and 0.◦4 for flights 1 and 2, re-
spectively, which are likely caused by altitude changes of
the balloon. In addition, pendulation modes (primarily at
15 s) were excited on occasion to amplitudes of 0.◦1–0.◦2,
with damping times of 10 min to half amplitude. All of
these modes were tracked with the camera in azimuth and
elevation and are included in the pointing solution. Their
effect on telescope roll is of second order and could not be
tracked.
3.2. Phases
The QMAP data set consists of three parallel data
streams:
radiometry data, chopper position, and CCD
star position. The optimum phases to associate a radiom-
etry sample with a chopper and a pointing position are
determined by maximizing the signal from Cas A with a
two-dimensional search in relative phase of the radiome-
try to both the chopper and the pointing. The two levels
of pointing modulation (chop and wobble) are essential to
breaking the 180◦ degeneracy encountered with one mod-
ulation only.
The phase fits for flight 2 were more complicated than
those for flight 1 because of the buffer overflows described
above. While the flight 1 chopper-to-pointing lag was
66 ms and the data-to-chopper lag was 46–48 ms, the cor-
responding lags for flight 2 were 12.17 s and 19.34–19.40 s,
respectively. Despite being large, these phases are accu-
rate to one chopper cycle for pointing and to 0.08 samples
(0.1 ms) for the data, indicating that there is no significant
phase smearing in either flight. The calibration pulses gave
an independent check of the phase solutions in both flights.
3.3. Gain Drift Correction
As described in D98, we inject calibration pulses of con-
stant strength and roughly 40 samples duration into each
radiometry channel before the first amplifier, allowing us
to monitor the total gain of each channel throughout the
flight. The pulses reveal that in both flights the instrument
gain drifted smoothly by about 10–15% in each channel.
These drifts were modeled and applied as corrections in
the subsequent analysis. The corrected gains vary < 1%
throughout the flight.
3.4. Radiometry Offsets
The chop-synchronous radiometry offset was determined
during mapmaking as described in dO98. This offset is
produced by the variation in effective emissivity of the
chopper plate (Wollack et al. 1997) and by the modula-
tion of emissive cavities in the optical layout. In the map-
ping analysis, the radiometry offset was calculated over
one chopper cycle without regard to the actual chopper
position. When the results are folded by chopper deflec-
tion, the offsets at overlapping positions are in agreement,
as illustrated in Figure 1 for the Ka1 channel. This vali-
dates both the phasing and the radiometry offset analysis.
Fig. 1.- Chop-synchronous radiometry offset for channel Ka1
from flight 1 (wide curve) and flight 2 (narrow curve) as a func-
tion of chopper deflection. Note the agreement of the folded curves,
which is similar in the other channels.
3.5. Beams
The beam patterns were measured as part of our system
calibration, with θc = 3◦ and ∆az = 5◦ in flight 1 and 1.◦4
in flight 2. Because the elevation of the beams is fixed, the
vertical component of the beam map is obtained by a drift
scan.
The beam maps were calculated by gridding the raw
data into 0.◦1 pixels. For flight 1, the chop-synchronous
radiometry offset discussed in section 3.4 was subtracted
directly from each channel, flattening the map background
Herbig et al.
3
by more than a factor of 10. In flight 2, the chop ampli-
tude for CMB observations was smaller than that used
for Cas A, so that the radiometry offset could not be sub-
tracted from the beam maps. Figure 2 shows two beam
images from flight 1. The maps for flight 2 are similar,
except for slightly larger background variations.
Fig. 2.- Flight 1 beam images of Cas A from the combined
Ka1/2 channels (left) Q1/2 channels (right). The chop-synchronous
radiometry offset was subtracted. The grey-scale values stretch lin-
early from −10 to +15 mK, and the axes are relative elevation and
azimuth in degrees on the sky. The structure to the bottom right
of Cas A corresponds to radio sources in the Galaxy.
For the beam shape determination, corresponding pairs
of beams (both polarizations for each feed, e.g., Ka1/2)
were combined and fitted as elliptical Gaussians. The
five shape parameters (az/el position, major and minor
FWHM, and angle) for the combined beams were then
used as constraints for the peak fits in each individual
channel. The beam sizes are 0.◦89 ± 0.◦03, 0.◦66 ± 0.◦02 and
0.◦70 ± 0.◦02 for Ka1/2, Q1/2 and Q3/4, respectively. The
robustness of these fits was explored by varying the fit
window size and the background subtraction model. The
resulting systematic error estimates in the beam integrals
(≈ 4–8%) exceed the formal errors on the fit parameters
and have been included in the overall calibration errors.
Our susceptibility to errors in the beam sizes is small:
for the calibration, only the error in the beam integral
is relevant. With our mapping approach, errors in the
beam size only affect the highest multipoles of the angular
power spectrum, as discussed in dO98. The beam param-
eters were measured independently for each flight and are
in good agreement.
3.6. Calibration
Our primary calibration source is the supernova rem-
nant Cas A, unresolved at our beam sizes, for which we
used the flux densities compiled by Baars et al. (1977)
augmented by data points between 31 and 250 GHz (Leitch
1998, Chini et al. 1984, Mezger et al. 1986). The result-
ing spectrum fit is log(Sν /Jy) = (5.713 ± 0.023) − (0.759 ±
0.006) log(ν/MHz) at epoch 1980.0. We use the secular de-
crease model from Baars et al. (1977) to predict the flux
density at our observing epochs. The uncertainty in the
flux density of Cas A is approximately 8.7% (including the
uncertainty in our center frequencies of ≈ 0.4 GHz) and
is the largest single contributor to the overall calibration
uncertainty.
Using the beam fits described in section 3.5, we predict
an antenna temperature for Cas A between 15 and 25 mK,
depending on band and channel. Each flight is calibrated
independently.
are automatically corrected to the appropriate calibration.
The total calibration uncertainties are between 10% and
13%, depending on channel and flight. Roughly equal
amounts are uncorrelated (beam fit and center frequency)
and correlated (Cas A) between channels.
3.7. Data Rejection
A small portion of the data (7.8% and 4.0% for flights 1
and 2, respectively) were rejected from the analysis. The
bulk (4.8% and 4.0%) were cut around the calibration
pulses (every 100 s) to allow recovery of the AC-coupled ra-
diometer baseline. Smaller cut windows resulted in a time-
dependent radiometry offset, while larger windows affected
neither the radiometry offset nor the final results. The re-
maining rejections were due to pointing glitches (0.7% and
0.02%) and to lost guide stars (2.3% in flight 1). Apart
from removing Q1 of flight 1 from analysis (see D98), there
were no cuts based on the contents or quality of the ra-
diometry data themselves.
4. RESULTS AND DISCUSSION
The techniques for producing sky maps and determin-
ing the angular power spectrum are described in detail in
dO98. The mapmaking step filters the data at the fun-
damental and the first harmonic of the chop frequency to
remove the chop-synchronous offset and pre-whitens the
time power spectrum to minimize the noise correlation be-
tween samples. The map is then derived from the linear
inversion of a highly over-determined system of equations
involving the filtered data points and their corresponding
sky pixels and residual scan-synchronous offsets. The final
map product consists of a vector x of N pixels (each with
some position on the sky) and its N × N noise covariance
matrix Σ.
Fig. 3.- Wiener-filtered maps from flight 2. The combined
Ka-band map (left), the combined Q-band map (middle) and the
combined map from all channels (right) are shown with the NCP is
at the center of the arcs and ra=0 at the top, increasing clockwise.
The rightmost map subtends 15◦
×16◦. These data are independent
of those in D98.
For visualization, Figure 3 shows the Wiener-filtered
versions of the maps (xw ≡ S[S + Σ]−1x), where S is
the covariance matrix due to the observed level of sky
fluctuations. Such filtering accentuates the statistically
significant features of a map.
The angular power spectrum of the sky in Figure 4
is produced directly from the combined unfiltered map
in each band by expanding in signal-to-noise eigenmodes
and sorting by decreasing signal-to-noise ratio. Ranges of
eigenmodes are averaged to produce the points shown in
Figure 4 and listed in Table 1. They do not contain the
calibration error of 12% for Ka and 11% for Q 1; this er-
1At the highest ℓ, the fraction of the calibration error due to uncertainty in the beam is partially offset by a change in the window function.
Note that the beams used for the calibration were ob-
served with the same strategy as the CMB data; this
means that most stable systematic errors that could lead
to defocussing (e.g., pointing noise and drifts, pendula-
tion, phases, and CCD camera and chopper calibrations)
To be conservative, we do not include this effect
4
QMAP 2
ror is systematic and is correlated between the bands of
QMAP as well as the Saskatoon results (Netterfield et al.
1997).
Fig. 4.- Angular power spectrum from flight 2. The central
QMAP point represents Q-band. The other two represent Ka-band
and have uncorrelated errors. The error bars include sample vari-
ance, but not the calibration error. For comparison, we show the
results from COBE and Saskatoon (see Tegmark 1996).
Checks for systematic errors can be made by differenc-
ing the maps of each polarization pair, which cover the
same sky pixels because they use the same feed. We use a
generalized χ2 statistic (dO98) to measure the signal-to-
noise ratio ν in each difference map xd = xa − rxb, where
a and b denote the two maps (e.g., Ka1/2). When r = 1,
the sky signal is subtracted completely and ν should be
a minimum. On the other hand, with r ≪ 1 (r ≫ 1),
map a (map b) dominates and ν gives the significance of
the anisotropy detection in this map. Figure 5 shows this
comparison for all three map pairs and illustrates that de-
spite the substantial signal (which is detected at the sig-
nificance level of 3σ, 2σ, 1σ, 6σ, 13σ and 20σ in the Ka1,
Ka2, Q1, Q2, Q3 and Q4 channels, respectively), none of
the difference maps show any significant remaining struc-
ture. Q3 and Q4 are particularly significant since they are
the only maps to cover the big cold spot seen in Figure 3
and in the Saskatoon data (Netterfield et al. 1997). The
spectral analysis of the data will be presented in a future
paper.
5. CONCLUSIONS
The second flight of QMAP produced a map of the
sky covering 83 square degrees at resolutions of 0.◦7–0.◦9
and frequencies of 42–31 GHz with ∼> 20σ detection of
anisotropy. Systematic checks strongly suggest that this
anisotropy is on the sky rather than produced by the in-
strument. We also present an angular power spectrum
covering 40 < ℓ < 200 produced from the map.
In this
region of the sky, the CMB is the dominant source of
anisotropy at 30-40 GHz (de Oliveira-Costa et al. 1997);
thus we interpret the observed signal as predominantly
CMB anisotropy. The raw data will be made public upon
publication of this Letter.
Table 1. - Flight 2 Angular Power Spectrum. The band powers
δTℓ = [ℓ(ℓ+1)Cℓ/2π]1/2 have window functions whose mean and rms
width are given by ℓef f and ∆ℓ. The errors δT include both noise
and sample variance, and are uncorrelated for the two Ka-points.
Band
Ka
Ka
Q
ℓef f ∆ℓ
47
64
67
91
145
125
δT [µK]
46+10
−12
63+10
−12
56+5
−6
s
n
o
i
t
a
i
v
e
d
d
r
a
d
n
a
t
S
20
15
10
5
0
4
r Q
3 -
Q
Q 1 -
r Q 2
r K a 2
K a 1 -
P
r
o
b
a
b
i
l
i
t
y
f
o
r
p
u
r
e
n
o
i
s
e
0.6
0.4
0.2
0.2
0.1
1
10
Relative normalization r
0.1
1
10
Relative normalization r
Fig. 5.- Evidence for real sky signal in the maps. The left panel
shows the signal-to-noise ratio ν in difference maps weighted by r
(see text) for Ka1 vs. Ka2 (solid black), Q1 vs. Q2 (dashed) and Q3
vs. Q4 (grey). The right panel shows the corresponding probability
that the difference map contains noise only. This figure shows that
all maps except Q1 contain significant detection of anisotropy and
that the signal is common to both maps of each polarization pair.
We are grateful for the invaluable support of the NSBF
and we thank Eric Torbet for his contributions. This
work was supported by a David & Lucile Packard Foun-
dation Fellowship (to LP), a Cottrell Award from Re-
search Corporation, an NSF NYI award, NSF grants
PHY-9222952 and PHY-9600015, NASA grant NAG5-
6034 and Hubble Fellowships HF-01044.01−93A (to TH)
and HF-01084.01−96A (to MT) from by STScI, operated
by AURA, Inc. under NASA contract NAS5-26555.
REFERENCES
Baars, J. W. M., Genzel, R., Pauliny-Toth, I. I. K., & Witzel, A.
1977, A&A, 61, 99
Chini, R., Kreysa, E., Mezger, P. G., & Gemund, H.-P. 1984, A&A,
137, 117
de Oliveira-Costa, A., Kogut, A., Devlin, M. J., Netterfield, C. B.,
Page, L., & Wollack, E. J. 1997, ApJ, 482, L17
de Oliveira-Costa, A., Devlin, M. J., Herbig, T., Miller, A.
D., Netterfield, C. B., Page, L., & Tegmark, M. 1998, astro-
ph/9808045 (dO98)
Devlin, M. J., de Oliveira-Costa, A., Herbig, T., Miller, A.
D., Netterfield, C. B., Page, L., & Tegmark, M. 1998, astro-
ph/9808043 (D98)
Leitch, E. M. 1998, Ph.D. Thesis, California Institute of Technology
Mezger, P. G., Tuffs, R. J., Chini, R., Kreysa, E., & Gemund, H.-P.
1986, A&A, 167, 145
Netterfield, C. B., Devlin, M. J., Jarosik, N., Page, L., & Wollack,
E. J. 1997, ApJ, 474, 47
Tegmark, M. 1996, Phys. Rev. D, 55, 5895
Wollack, E. J., Devlin, M. J., Jarosik, N., Netterfield, C. B., Page,
L., & Wilkinson, D. 1997, ApJ, 476, 440
n
|
0806.0850 | 1 | 0806 | 2008-06-04T20:06:34 | Substructure and Scatter in the Mass-Temperature Relations of Simulated Clusters | [
"astro-ph"
] | Galaxy clusters exhibit regular scaling relations among their bulk properties. These relations establish vital links between halo mass and cluster observables. Precision cosmology studies that depend on these links benefit from a better understanding of scatter in the mass-observable scaling relations. Here we study the role of merger processes in introducing scatter into the $M$-$T_{\rm X}$ relation, using a sample of 121 galaxy clusters simulated with radiative cooling and supernova feedback, along with three statistics previously proposed to measure X-ray surface brightness substructure. These are the centroid variation ($w$), the axial ratio ($\eta$), and the power ratios ($P_{20}$ and $P_{30}$). We find that in this set of simulated clusters, each substructure measure is correlated with a cluster's departures $\delta \ln T_{\rm X}$ and $\delta \ln M$ from the mean $M$-$T_{\rm X}$ relation, both for emission-weighted temperatures $T_{\rm EW}$ and for spectroscopic-like temperatures $T_{\rm SL}$, in the sense that clusters with more substructure tend to be cooler at a given halo mass. In all cases, a three-parameter fit to the $M$-$T_{\rm X}$ relation that includes substructure information has less scatter than a two-parameter fit to the basic $M$-$T_{\rm X}$ relation. | astro-ph | astro-ph | Draft version November 29, 2018
Preprint typeset using LATEX style emulateapj v. 03/07/07
8
0
0
2
n
u
J
4
]
h
p
-
o
r
t
s
a
[
1
v
0
5
8
0
.
6
0
8
0
:
v
i
X
r
a
SUBSTRUCTURE AND SCATTER IN THE MASS -- TEMPERATURE RELATIONS OF SIMULATED CLUSTERS
David A. Ventimiglia1, G. Mark Voit1, Megan Donahue1, S. Ameglio2
Draft version November 29, 2018
ABSTRACT
Galaxy clusters exhibit regular scaling relations among their bulk properties. These relations estab-
lish vital links between halo mass and cluster observables. Precision cosmology studies that depend
on these links benefit from a better understanding of scatter in the mass-observable scaling relations.
Here we study the role of merger processes in introducing scatter into the M -TX relation, using a
sample of 121 galaxy clusters simulated with radiative cooling and supernova feedback, along with
three statistics previously proposed to measure X-ray surface brightness substructure. These are the
centroid variation (w), the axial ratio (η), and the power ratios (P20 and P30). We find that in this
set of simulated clusters, each substructure measure is correlated with a cluster's departures δ ln TX
and δ ln M from the mean M -TX relation, both for emission-weighted temperatures TEW and for
spectroscopic-like temperatures TSL, in the sense that clusters with more substructure tend to be
cooler at a given halo mass. In all cases, a three-parameter fit to the M -TX relation that includes
substructure information has less scatter than a two-parameter fit to the basic M -TX relation.
Subject headings: galaxies: clusters: general, X-rays: galaxies: clusters
1. INTRODUCTION
Clusters of galaxies play a critical role in our under-
standing of the Universe and its history and are po-
tentially powerful tools for conducting precision cosmol-
ogy. For example, large cluster surveys can discrimi-
nate between cosmological models with different dark-
energy equations of state by providing complementary
observations of the shape of the cluster mass function,
evolution in the number density of clusters with red-
shift, and bias in the spatial distribution of clusters
(Wang & Steinhardt 1998; Levine et al. 2002; Hu 2003;
Majumdar & Mohr 2004; Voit 2005). However, this po-
tential to put tight constraints on the properties of dark
energy will be realized only if we can accurately measure
the masses of clusters and can precisely characterize the
scatter in our mass measurements.
Scatter in X-ray cluster properties is directly related to
substructure in the intracluster medium. If clusters were
all structurally similar, then there would be a one-to-one
relationship between halo mass and any given observable
property. Generally speaking, deviations from a mean
mass-observable relationship are attributed to structural
differences among clusters. One kind of structural differ-
ence is the presence or absence of a cool core, in which
the central cooling time is less than the Hubble time at
the cluster's redshift, and the prominence of a cool core is
observed to be a source of scatter in scaling relationships
(Fabian et al. 1994; Markevitch 1998; Voit et al. 2002;
McCarthy et al. 2004). We also expect structural dif-
ferences to arise from substructure in the dark matter,
galaxy, and gas distributions. For instance, there may be
a spread in halo concentration at a given mass, variations
in the incidence of gas clumps, differences in the level of
AGN feedback, or various effects due to mergers. All of
1 Michigan State University, Physics & Astronomy Dept., East
Lansing, MI 48824-2320; [email protected], [email protected],
[email protected]
2 Dipartimento di Astronomia dell'Universit`a di Trieste, via
Tiepolo 11, I-34131 Trieste, Italy; [email protected]
these deviations can be considered forms of substructure
that produce scatter in the mass-observable relations one
would like to use for cosmological purposes. While it may
ultimately be possible to constrain the amount of scat-
ter and its evolution with redshift using self-calibration
techniques (Lima & Hu 2005), such constraints would be
improved by prior knowledge about the relationship be-
tween scatter and substructure.
Traditionally, the most worrisome form of substructure
has been that due to the effects of merger events. Clus-
ters are often identified as "relaxed" or "unrelaxed", with
the former assumed to be nearly in hydrostatic equilib-
rium and the latter suspected of being far from equilib-
rium. Cosmological simulations of clusters indicate that
the truth is somewhere in between. The cluster popula-
tion as a whole appears to follow well-defined virial rela-
tions with log-normal scatter around the mean, showing
that clusters do not cleanly separate into relaxed and
unrelaxed systems (e.g.,Evrard et al. (2007)). Even the
most relaxed-looking clusters are not quite in hydrostatic
equilibrium (e.g.,Kravtsov et al. (2006)). Instead of sim-
ply being "relaxed" or "unrelaxed," clusters occupy a
continuum of relaxation levels determined by their re-
cent mass-accretion history.
Quantifying this continuum of relaxation offers oppor-
tunities for reducing scatter in the mass-observable rela-
tions. If mergers are indeed responsible for much of the
observed scatter around a given scaling relation, then
there may be correlations between a cluster's morphol-
ogy and its degree of deviation from the mean relation.
Once one identifies a morphological parameter that cor-
relates with the degree of deviation, one can construct a
new mass-observable relation with less scatter by includ-
ing the morphological parameter in the relation. Such
an approach would be analogous to the improvement of
Type Ia supernovae as distance indicators by using light-
curve stretch as a second parameter to indicate the su-
pernova's luminosity (Phillips 1993; Riess et al. 1996).
Here we investigate how merger-related substructure
in simulated clusters affects the relationship between a
2
simulated cluster's mass and the temperature of its in-
tracluster medium, building upon Buote & Tsai (1995)
and O'Hara et al. (2006). Buote & Tsai (1995) quanti-
fied the morphologies and dynamical states of observed
clusters and found structure measures to be an indica-
tor of the dynamical state of a cluster. O'Hara et al.
(2006) also examined morphological measurements, for
both observed and simulated clusters, and found that
simulations without cooling showed no correlation be-
tween substructure and scaling relation scatter. In this
work we examine substructure for simulated clusters with
radiative cooling and focus on the idea that merger pro-
cesses introduce intrinsic scatter into the M -TX relation-
ship by displacing clusters in the M -TX plane away from
the mean X-ray temperature hTXiM at a given mass
M , either to higher or lower average ICM temperature.
We then adopt a set of statistics (Buote & Tsai 1995;
O'Hara et al. 2006) for quantifying galaxy cluster sub-
structure and merger activity in order to investigate this
hypothesis. Section 2 discusses the M -TX scaling rela-
tionship in our sample of simulated clusters and shows
that disrupted-looking clusters in this sample tend to
be cooler at a given cluster mass.
In Section 3 we at-
tempt to quantify the relationship between morphology
and temperature using four different substructure statis-
tics and compare it to similar studies. We then show
that substructure in these simulated clusters indeed cor-
relates with scatter in the M -TX relationship and assess
the prospects for using that correlation to reduce scatter
in the M -TX plane. Section 4 summarizes our results.
2. MASS-TEMPERATURE RELATION IN SIMULATED
CLUSTERS
This study is based on an analysis of 121 clus-
ters simulated using the cosmological hydrodynamics
TREE+SPH code GADGET-2 (Springel 2005), which
were simulated in a standard Λ cold dark matter
(ΛCDM) universe with matter density ΩM = 0.3, h =
0.7, Ωb = 0.04, and σ8 = 0.8. The simulation treats ra-
diative cooling with an optically-thin gas of primordial
composition, includes a time-dependent UV background
from a population of quasars, and handles star forma-
tion and supernova feedback using a two-phase fluid
model with cold star-forming clouds embedded in a hot
medium. All but four of the clusters are from the simu-
lation described in Borgani et al. (2004), who simulated
a box 192 h−1 Mpc on a side, with 4803 dark matter par-
ticles and an equal number of gas particles. The present
analysis considers the 117 most massive clusters within
this box at z = 0, which all have M200 greater than
5 × 1013 h−1M⊙. By convention, M∆ refers to the mass
contained in a sphere which has a mean density of ∆
times the critical density ρc, and whose radius is denoted
by R∆
That cluster set covers the ∼1.5-5 keV temperature
range, but the 192 h−1 Mpc box is too small to contain
significantly hotter clusters. We therefore supplemented
it with four clusters with masses > 1015 h−1 M⊙ and tem-
peratures > 5 keV drawn from a dark-matter-only sim-
ulation in a larger 479 h−1 Mpc box (Saro et al. 2006).
The cosmology for this simulation also was ΛCDM, but
with σ8 = 0.9. These were then re-simulated including
hydrodynamics, radiative cooling, and star formation,
again with GADGET-2 and using the zoomed-initial-
conditions technique of Tormen (1997), with a fourfold
increase in resolution. This is comparable to the reso-
lution of the clusters in the smaller box. Adding these
four massive clusters to our sample gives a total of 121
clusters with M200 in the interval 5 × 1013 h−1M⊙ to
2 × 1015 h−1M⊙.
We first need to specify our definitions for mass and
temperature. In this paper, cluster mass refers to M200.
For temperature, we use two definitions. The first is the
emission-weighted temperature
TEW = R T [n2Λ(T )] d3x
R n2Λ(T ) d3x
,
(1)
where n is the electron number density and Λ(T )
is the usual cooling function for intracluster plasma.
The second is the spectroscopic-like temperature of
Mazzotta et al. (2004),
TSL = R T [n2T −3/4] d3x
R n2T −3/4 d3x
,
(2)
where the power-law weighting function replacing Λ(T )
is chosen so that TSL approximates as closely as possible
the temperature that would be determined from fitting
a single-temperature plasma emission model to the inte-
grated spectrum of the intracluster medium. The pres-
ence of metals in the ICM of real clusters introduces line
emission that complicates the computation of TSL for
clusters <3 keV (Vikhlinin 2006). However, the simu-
lated spectra for the clusters in our sample are modelled
with zero metallicity, which eases this restriction in our
analysis.
Figure 1 shows the mass-temperature relations based
on these definitions for our sample of simulated clusters.
The best fits to the power-law form
M = M0(cid:18) TX
3 keV(cid:19)α
(3)
have the coefficients M0 ≃ 5.9 × 1013 h−1M⊙, α ≃
1.5 for TX corresponding to TSL and M0 ≃ 4.4 ×
1013 h−1M⊙, α ≃ 1.4 for TX corresponding to TEW. As is
generally the case for simulated clusters, the power-law
indices of the mass-temperature relations found here are
consistent with cluster self-similarity and the virial the-
orem (Kaiser 1986; Navarro et al. 1995). These relation-
ships have scatter, which we characterize by the standard
deviation in log space σln M about the best-fit mass at
fixed temperature TX. When relating M to the emission-
weighted temperature TEW, we find σln M = 0.102.
When relating cluster mass M to the spectroscopic-like
temperature TSL, the scatter is σln M = 0.127. That the
scatter is larger for the spectroscopic-like temperature is
not surprising, given the sensitivity of TSL to the thermal
complexity of the ICM.
Figure 1 also highlights two subsamples for each defi-
nition of temperature, selected based on the clusters' de-
viations in ln TX space from the mean mass-temperature
relation. In each panel, open circles represent the eight
clusters that have the largest positive deviations and are
therefore "hotter" than other clusters of the same mass,
while filled circles represent the eight with the largest
negative deviation and are "cooler" than other clusters of
the same mass. In general, these temperature estimates
3
Fig. 1. -- Mass-temperature (M -TX) relationships for the 121
clusters in our sample. Spectroscopic-like temperature TSL is plot-
ted in the top panel, while emission-weighted temperature TEW
is plotted in the bottom panel. The solid lines show the best-fit
power-law relation M = M0(TX/3 keV)α over the whole sample,
while the dashed lines show the best-fit relation for systems with
TX > 2 keV. Open circles represent the clusters that have the
greatest positive temperature offset from the mean relationship,
and filled circles represent the clusters with the greatest negative
temperature offset.
are well correlated, so that hotter outliers in TEW are
also hotter outliers in TSL and cooler outliers in TEW are
also cooler outliers in TSL. Since Figure 1 distinguishes
the most extreme outliers for the two temperature esti-
mates, this distinction may define slightly different sets,
though they still overlap.
Figure 2 presents a gallery of surface brightness maps
for two sets of eight clusters with the most extreme off-
sets from the mean M -TSL relation. The eight unusually
hot clusters are in the top panel, and the eight cooler
clusters are in the bottom panel. In these plots the hot-
ter clusters appear more symmetric, and are seemingly
"more relaxed," and the cooler clusters appear less sym-
metric and seemingly "less relaxed." The gallery as a
whole therefore suggests that relaxed clusters tend to be
hot for their mass and unrelaxed clusters tend to be cool
for their mass.
At first glance, the result that disrupted-looking clus-
ters in cosmological simulations tend to be cooler than
other clusters of the same mass may seem counterintu-
itive, since one might expect that mergers ought to pro-
duce shocks that raise the mean temperature of the in-
tracluster medium. This finding has also been noted by
Mathiesen & Evrard (2001) and Kravtsov et al. (2006).
Cluster systems in the process of merging tend to be cool
for their total halo mass because much of the kinetic en-
ergy of the merger has not yet been thermalized.
The idealized simulations of Poole et al. (2007) illus-
trate what may happen to the ICM temperature during
a single merger. Before the cores of the two merging sys-
tems collide, the mean temperature is cool for the over-
Fig. 2. -- Surface-brightness contour maps for sixteen of the clus-
ters in our sample, overlaid with a circular aperture of radius R500.
The top panel has eight maps representing the clusters that have
the largest positive deviation δ ln TX from the mean M -TX relation,
calculated using the spectroscopic-like temperature TSL. These are
the objects represented by open circles in the top panel of Figure
1. The bottom panel has eight maps representing the clusters that
have the largest negative deviation from the same M -TX relation-
ship, represented by the filled circles in Figure 1. Note that the
"hotter" clusters in the top panel appear more symmetric, while
the "cooler" clusters in the bottom panel appear more irregular.
all halo mass because it is still approximately equal to
the pre-merger temperature of the two individual merg-
ing halos. There is a brief upward spike in temperature
when the cores of the merging halos collide, after which
the system is again cool for its mass. Then, as the re-
maining kinetic energy of the merger thermalizes over
a period of a few billion years, the temperature gradu-
ally rises to its equilibrium value. The merging system
therefore spends a considerably longer time at relatively
cool values of mean temperature for its halo mass than
at relatively hot values. Hence, such simulations sug-
gest a possible explanation for why more relaxed sys-
tems would tend to lie on the hot side of the M -TX
relation, while disrupted systems would tend to lie on
the cool side. A caveat, however, is that the current
generation of hydrodynamic cluster simulations tend to
produce relaxed clusters whose temperature profiles con-
tinue to rise to smaller radii than is observed in real
clusters (Tornatore et al. 2003; Nagai et al. 2007), po-
tentially enhancing average temperatures for such sys-
tems. As a separate test of this effect, we excise the core
regions from our sample clusters, calculate new substruc-
ture measures and new emission-weighted temperatures
for the core-excised clusters, and repeat our analysis.
3. QUANTIFYING SUBSTRUCTURE
The question we would like to address in this study
is whether the surface-brightness substructure evident
in Figure 2 is well enough correlated with deviations
from the mean mass-temperature relation to yield use-
ful corrections to that relation. In order to answer that
question, we need to quantify the surface-brightness sub-
4
structure in each cluster image, so that we can deter-
mine the degree of correlation across the entire sample.
O'Hara et al. (2006) explored the relationship between
cluster structure and X-ray scaling relations in both ob-
served and simulated clusters, and we adopt their suite
of substructure measures in this study. These include
centroid variation, axial ratio, and the power ratios of
Buote & Tsai (1995). In this section we define and dis-
cuss those statistics and apply them to surface-brightness
maps made from three orthogonal projections of each
cluster. Then we assess how well these statistics correlate
with offsets from the mean mass-temperature relation.
3.1. Axial Ratio
The axial ratio η for a cluster surface-brightness map is
a measure of its elongation, which is of interest because
it has been found from simulations that the ICM is of-
ten highly elongated during merger events (Evrard et al.
1993; Pearce et al. 1994). It is computed from the second
moments of the surface brightness,
Mij =X IXxixj .
The summation is conducted over the coordinates
(x1, x2) of the pixels that lie within an aperture centered
at the origin of the coordinate system to which (x1, x2)
refer. Following the work of O'Hara et al. (2006), we use
an aperture of radius R500 centered on the brightness
peak. We then compute η from the ratio of the non-zero
elements that result from diagonalizing the matrix M .
That is,
(4)
Fig. 3. -- Surface-brightness contour plot of an asymmetric clus-
ter which, for certain choices of aperture placement, yields an axial
ratio close to 1. The circle represents an aperture of R500
.
D = U T M U,
where U is a diagonalizing matrix for M , and
η =( D12
D21
D21
D12
, D12 ≤ D21
, D12 > D21 ) .
(5)
(6)
The axial ratio is therefore defined to be in the range
η ∈ [0, 1], with η = 1 for a circular cluster. Of course
there are other choices for the origin of the coordinate
system, besides using the brightness peak. For instance,
in order to avoid misplaced apertures yielding artifi-
cially low axial ratios for nearly circular distributions,
one could adjust the position of the aperture to seek a
maximum in η. Doing this, we sometimes find that η ≈ 1
even for non-circular clusters, as is evident in Figure 3.
This figure depicts the surface-brightness map of what
appears to be a disturbed cluster, chosen from among
those in our sample that appear by eye to be the most
unrelaxed. Yet, it happens to have an axial ratio very
close to 1 for an aperture placed so as to maximize η.
This example demonstrates that, while the axial ratio
statistic may yield results consistent with a visual in-
terpretation of cluster substructure, it is also capable of
unexpected results for some clusters.
To further illustrate this point, we have computed an
axial ratio value for this cluster for every possible choice
of aperture placement. Apertures of radius R500 were
centered on each and every pixel within the surface-
brightness map, provided the aperture so placed does
not reach the edge of the map. This procedure gener-
ated an axial-ratio "surface" mapping all of the aperture
placements to a value of η. Figure 4 shows the axial-ratio
Fig. 4. -- Surface of axial ratio η as a two-dimensional function
of the coordinates of the aperture center. The axial ratio statistic
appears to be ill-defined for this cluster.
Fig. 5. -- Abstract surface of axial ratio η as a two-dimensional
function of the coordinates of the aperture center. This is a relaxed
cluster, for which the axial ratio is better-defined.
surface for the cluster in Figure 3. For comparison pur-
poses, Figure 5 presents an axial-ratio surface map for
a very symmetric, uniform, and apparently relaxed clus-
ter, in which the cluster's brightness peak reassuringly
corresponds to the aperture location that maximizes η.
In contrast, the presence of two peaks in the axial-ratio
surface for the asymmetric cluster shows that η can some-
times depend strongly on aperture placement. Ideally, we
would like to place the aperture on the "center" of this
5
Fig. 6. -- Relationship between a cluster's deviation δ ln TX from the mean M -TX relationship and four measures of substructure: the
centroid variation w, the axial ratio η, and the power ratios P20 and P30. The lower panels are for spectroscopic-like temperature, and
the upper ones are for emission-weighted temperature. The light gray bands and vertical dashed lines show the extremes in substructure
measurements, with 80% of our clusters falling in the central region between the bands. The black filled circles correspond to clusters above
2 keV, while the plus signs correspond to clusters below 2 keV. Finally, the solid lines indicate the best-fitting linear relationships between
our substructure measures and temperature offset, for the above 2 keV sub-sample denoted by the black filled circles.
cluster, but the center of an unrelaxed cluster can be
difficult to define, meaning that the axial ratio statistic
may be likewise ill-defined for such clusters.
3.2. Power Ratio
power-ratio
statistics
The
(Buote & Tsai
1995;
O'Hara et al. 2006) quantify substructure by decompos-
ing the surface-brightness image into a two-dimensional
multipole expansion, the terms of which are calculated
from the moments of the image, computed within an
aperture of radius Rap:
am(Rap) =ZR′≤Rap
bm(Rap) =ZR′≤Rap
Σ(~x′)(R′)m cos mφ′ d2x′
(7)
Σ(~x′)(R′)m sin mφ′ d2x′.
(8)
The power in terms of order m is then
3.3. Centroid Variation
The centroid variation statistic w is a measure of the
center shift, or "skewness", of a two-dimensional pho-
ton distribution.
It is measured for a cluster surface-
brightness map in the following way. For a set of surface-
brightness levels one finds the centroids of the corre-
sponding isophotal contours and computes the variance
in the coordinates of those centroids, scaled to R500. Here
we select 10 isophotes evenly spaced in log IX between
the minimum and maximum of IX within an aperture
of radius R500 centered on the brightness peak, so as
to adapt to the full dynamic range of surface bright-
ness for different clusters. We employed this adaptive
scheme because using one set of isophotes for all clusters
tended to ignore important substructure in less massive
clusters when they had surface brightness substructure
inside R500 but outside of the lowest isophote.
Pm =
(a2
m + b2
m)
2m2R2m
ap
.
(9)
3.4. Substructure and Scaling Relationships
For m = 0, the power is given by
P0 = [a0 ln(Rap)]2.
(10)
The power ratios Pm0 ≡ Pm/P0 are then dimensionless
measures of substructure which have differing interpreta-
tions. For instance, P10 quantifies the degree of balance
about some origin and can be used to find the image cen-
troid, P20 is related to the ellipticity of the image, and
P30 is related to the triangularity of the photon distri-
bution. As in the case of the axial ratio computations,
we set the aperture radius Rap equal to R500. The most
appropriate place to center the aperture is at the set of
pixel coordinates that minimizes P10, which we achieve
using a self-annealing algorithm.
Using the quantitative measures of substructure de-
scribed in the previous section, we can test the signifi-
cance of the relationship between substructure and tem-
perature offset hinted at in Figure 2. We begin by treat-
ing four of our substructure statistics -- centroid variation
w, axial ratio η, and power ratios P20 and P30 -- as dif-
ferent imperfect measurements of an intrinsic degree of
substructure S. Figure 6 shows the relationship between
substructure and a cluster's deviation δ ln TX from the
mean M -TX relation for each substructure measure. In
each case we present results for both the spectroscopic-
like temperature TSL and the emission-weighted temper-
ature TEW. Note that centroid variation w and the power
ratios P20 and P30 have large dynamic ranges, whereas
6
the axial ratio η is always of order unity. We therefore
attempt to fit the relationships between δ ln TX and the
different substructure measures with the following forms:
Awα
B + βη
CP γ
20
DP λ
30
(11)
δ ln TX =
To visually indicate where the bulk of our substruc-
ture measures lie, Figure 6 has light gray bands cov-
ering the extremes, so that 80% of our sample clus-
ters have substructure measures lying between the ex-
tremes. The power ratios in our study generally span
two decades (in units of 10−7), from ∼2 -- 300 for P20
and from ∼0.01 -- 10 for P30. These ranges are consis-
tent with those of Buote & Tsai (1995), O'Hara et al.
(2006), and Jeltema et al. (2007). The measurements of
axial ratio in our sample, with 80% of clusters having
η∼0.4 -- 0.95, cover a slightly wider range than do the
simulated clusters of O'Hara et al. (2006). Finally, our
measurements of centroid variation, with 80% of clusters
having w[R500]∼0.01 -- 0.1, are again similar to those of
O'Hara et al. (2006).
As denoted in Figure 6 by black filled circles, the sys-
tems with TX above 2 keV occupy a slightly narrower
range of substructure values than the systems below 2
keV, which are denoted by plus signs. For the axial ra-
tio and the power ratios, the variance is 15 to 25 per-
cent larger among the low-temperature systems when
compared to the systems with TX > 2 keV. For cen-
troid variation the variance among the low-temperature
systems is approximately the same as it is among the
high-temperature systems. However, it is not clear that
there is a significant correlation between substructure
and mass, since the mean substructure values are gener-
ally very similar between the low-temperature and high-
temperature subsamples. The mean value of the power
ratio P30 is significantly larger for the low-temperature
subsample, however this measure also has the weakest
correlation with offsets from the mean M -TX relation.
To test whether the low-mass clusters in our sample
significantly boost the overall scatter in the M -TX rela-
tion, we perform a cut at 2 keV and fit this relation both
to the whole sample and to the sub-sample above 2 keV.
Figure 7 shows the residuals in mass, actual minus pre-
dicted, where the predicted mass derives only from the
M -TX relation. The plus signs indicate clusters whose
mass is predicted from an M -TX relation derived from
all 121 clusters. The black filled circles indicate clusters
that are above 2 keV in X-ray temperature, with the mass
estimated using the sub-sample M -TX relation. There is
a negligible reduction in scatter, from 0.127 to 0.124 for
TSL and from 0.102 to 0.094 for TEW, suggesting that at
best only a modest improvement is found in our sample if
we remove the low-mass systems. In order to test the de-
gree to which incorporating substructure measures adds
to this modest improvement, when we compare mass es-
timates derived using substructure to those derived only
from the M -TX relation, we focus on clusters above 2
keV in the rest of our analysis.
Figure 6 shows that for our simulated clusters, a
greater amount of measured substructure tends to be as-
sociated with "cooler" clusters while less substructure
tends to be associated with "hotter" clusters. Also, the
Fig. 7. -- Comparison of mass offset δ ln M between true mass
and predicted mass, based on the M (TX) relation. Plus signs indi-
cate residuals for masses estimated from the M -TX relation derived
from all 121 clusters, while black filled circles are for masses esti-
mated from the M -TX relation for clusters above 2 keV. Upper
panels are for spectroscopic-like temperature and lower panels are
for emission-weighted temperature. The standard deviations σ in
the residuals are given in each plot.
centroid variations w are more highly-correlated with
δ ln TX than are the other substructure parameters. We
interpret this to mean that the centroid variation is a
better predictor of the offset in the M -TX relationship
than are the power ratios and the axial ratio, though all
four measures appear to be related to the temperature
offset. Again, in this figure we denote systems above 2
keV by black filled circles, and systems below 2 keV by
plus signs.
Correlations between substructure and δ ln TX can po-
tentially be exploited to improve on mass estimates of
real clusters derived from the mass-temperature rela-
tion.
Instead of computing the temperature offset at
fixed mass, we can determine a substructure-dependent
mass offset at fixed temperature and then apply it as a
correction to the predicted mass Mpred(TX) one would
derive from the mean M -TX relation alone. To assess
the prospects for such a correction, based on this sample
of simulated clusters, we first define the mass offset from
the mean mass-temperature relation to be
δ ln M (TX) = ln(cid:20) M
Mpred(TX)(cid:21) ,
(12)
where M is the cluster's actual mass, and examine the
correlations between substructure measures and δ ln M .
Figure 8 shows the results. These plots show mass pre-
dictions from both the M -TSL relation and the M -TEW
relation. Consistent with our analysis of δ ln TX, the cen-
troid variation w appears to be a more effective predictor
of the mass offset δ ln M (TX). Nonetheless, all four mea-
sures of substructure appear to be correlated with mass
offset.
7
Fig. 8. -- Relationship between substructure and mass offset δ ln M (TX) from the mean M -TX relationship for the same substructure
measures as in Figure 6. The lower panels are for TX = TSL, and the upper panels are for TX = TEW. As in Figure 6 the solid lines indicate
the best-fitting linear relationships to the above 2 keV sub-sample denoted by black filled circles, and the gray bands and vertical dashed
lines mark the extremes in substructure between which 80% of our clusters lie. Also as in Figure 6, the plus signs correspond to systems
below 2 keV., while the black filled circles correspond to systems above 2 keV.
In order to incorporate a substructure correction into
the mass-temperature relation, we perform a multiple
regression, fitting our simulated clusters' mass, temper-
ature, and substructure data to the form,
log Mpred(TX, S) = log M0 + α log TX + βS,
(13)
where S represents one of the following substructure
measures: η, log w, log P20, or log P30. This fit gives us a
substructure-corrected mass prediction Mpred(TX, S) for
each substructure measure, and we can assess the effec-
tiveness of that correction by measuring the dispersion
of the substructure-corrected mass offset
δ ln M (TX, S) = ln(cid:20)
(14)
M
Mpred(TX, S)(cid:21) ,
between the revised prediction and the true cluster mass.
Figure 9 shows the results of that test. Open circles in
each panel indicate mass offsets δ ln M (TX) without sub-
structure corrections, which have a standard deviation
σM(T ). Filled circles indicate mass offsets δ ln M (TX, S)
with substructure corrections, which have a standard de-
viation σM(T,S). The upper set of panels shows results for
TSL, and the lower set is for TEW. In each case, incorpo-
rating a substructure correction to the mass-temperature
relation reduces the scatter, yielding more accurate mass
estimates. The centroid variation corrections are the
most effective, reducing the scatter in mass from 0.124
to 0.085 in the M -TSL relation and from 0.094 to 0.072
in the M -TEW relation, though admittedly this is again
a modest improvement. Although non-negligible struc-
ture correlates significantly with offsets in the M -TX
plane, apparently it does so with substantial scatter.
This scatter may be partly due to projection effects, in
which line-of-sight mergers are discounted by the mea-
sures of substructure and may dilute their corrective
power (Jeltema et al. 2007).
Lastly, Figure 10 shows the results for a similar analy-
sis to that of Figure 9, except that in this case we have
excised a region of radius 0.15R500 around the center of
each cluster and recomputed TEW. We do this to test
whether offset in temperature, whose correlation with
substructure is the basis of our correction scheme, stems
from a potentially unrealistic feature, which is that the
cores of many real clusters have temperature profiles that
decline at larger radii than occurs in simulated clusters.
As in Figure 9, we restrict our analysis to clusters above
2 keV. After doing this test, for TEW excising the core ac-
tually increases the scatter in M -TX from 0.094 to 0.106.
It may be that by removing the bright central region, the
average temperature becomes more sensitive to structure
outside the core. Also, this figure shows that the effect of
incorporating substructure measurements into the mass-
estimates is still present. The scatter is reduced to 0.075
for w, 0.093 for η, 0.090 for P20 and 0.094 for P30. Figure
10 summarizes the results of this test, which support the
conclusion that the reduction in scatter we realize using
substructure is a real effect and not an artifact of known
defects in the simulations.
3.5. Comparisons with Other Substructure Studies
O'Hara et al. (2006) examined the relationship be-
tween galaxy cluster substructure and X-ray scaling re-
lationships, including the M -TX relation, using both a
flux-limited sample of nearby clusters and a sample of
simulated clusters, and found a greater amount scatter
among the more relaxed clusters in their observed sam-
ple. Contrasting that result they also found a greater
amount of scatter among the more disrupted clusters in
their simulation sample, though they characterize the ev-
idence for this second result to be weak. Finally, they see
8
Fig. 9. -- Comparison of mass offset δ ln M between true mass and predicted mass, based on the M (TX) relation (open circles) and the
M (TX, S) relation (filled circles). Upper panels are for TSL and lower panels are for TEW. The standard deviations σ in the residuals are
given in each plot.
no evidence in either sample for more disrupted clusters
to be below the mean, and the more relaxed clusters to
be above. One difference between our study and theirs
is the presence of radiative cooling and supernova feed-
back in the simulation that produced our cluster sample.
Also, the focus of our work is different from theirs in that
we concentrate on the degree of correlation between the
amount of substructure and the size and direction of the
offset from the mean relation. We do find significant ev-
idence of this correlation, such that relaxed clusters are
hotter than expected given their mass. We also test, as
best we can given our simulation sample, the hypothesis
that substructure can be used to improve mass estimates
derived from the ICM X-ray temperature. It is possible
that our detection of a correlation between substructure
and temperature offset arises from the additional physics
in our simulated clusters, since when radiative cooling is
included, cool lumps may be better preserved than in
simulations that don't include cooling.
Our results are in agreement with Valdarnini (2006),
who examined substructure in clusters simulated with
cooling and feedback and found that unrelaxed clusters,
identified with a larger power ratio P30, have spectral-fit
temperatures biased low relative to the mass-weighted
temperatures. This trend aligns with our finding that
the spectroscopic-like temperature TSL is lower than the
best-fit temperature at fixed mass for clusters with larger
power ratios P20 and P30. However, Valdarnini (2006)
did not investigate the effectiveness of substructure mea-
sures in reducing scatter in the mass-temperature rela-
tion.
Our results are also in agreement with some of the
results of Jeltema et al. (2007), who have recently in-
vestigated correlations between substructure and offsets
in mass predictions in simulated clusters. They found
that measuring cluster structure is an effective way to
correct masses estimated using the assumption of hy-
drostatic equilibrium, which tend to be underestimates.
Our findings support these results, given that we find
substructure can be used to correct masses estimated di-
rectly from the M -TX relationship. There also are dif-
ferences between our findings and theirs. They report
that the M -TX relation for their simulation sample shows
no dependence on structure, whereas the clusters in our
sample exhibit offsets that correlate with the degree of
substructure. One possibility is that these differences
stem from differences in the simulations' feedback mech-
anisms. Another possibility is that some of the offset we
observe derives from enhanced temperatures in simula-
tions with radiative cooling. As we describe in section
3, we perform a test in which we estimate TEW using
projected surface-brightness and temperature maps, in
order to remove the core regions from our analysis, but
this may be less effective than properly excising the cores
in the simulations, as Jeltema et al. (2007) have done.
Kravtsov et al. (2006) also looked at the relationship
that cluster structure has to the M -TX relation in simu-
lated clusters, to show that the sensitivity of mass proxies
YX and YSZ to substructure is not very strong. They
divided their sample into unrelaxed and relaxed sub-
samples, based on the presence or absence of multiple
peaks in the surface-brightness maps of clusters, and
found the normalization of the M -TX relation to be bi-
ased to cooler temperatures for the unrelaxed systems.
Other workers also have looked at the relationship be-
tween the M -TX relation and substructure, as reflected
in the X-ray spectral properties. Mathiesen & Evrard
(2001) have examined the ratio of X-ray spectral-fit tem-
peratures in hard and full bandpasses for an ensemble of
simulated clusters, and found it to be a way of quanti-
fying the dynamical state of a cluster. We consider our
approach of using surface-brightness morphology infor-
9
Fig. 10. -- Comparison of mass offset δ ln M between true mass and predicted mass, based on the M (TX) relation (open circles) and
the M (TX, S) relation (filled circles). These results are for are for TEW, with the central 0.15 R500 region removed both from the average
temperature and from the substructure measures. The standard deviations σ in the residuals are given in each plot.
mation to be complementary to theirs. More recently,
Kay et al. (2007) performed an interesting analysis on
another large-volume simulation sample, using as sub-
structure metrics the centroid variation and measures
of concentration to report evolution in the luminosity-
temperature relationship. Specifically, they report that
the more irregular clusters in their sample lie above the
mean M -TX relation (i.e., they are cooler than average),
for the spectroscopic-like temperature TSL.
4. SUMMARY
Using a sample of galaxy clusters simulated with cool-
ing and feedback, we investigated three substructure
statistics and their correlations with temperature and
mass offsets from mean scaling relations in the M -TX
plane. First, we showed that the substructure statistics
w, η, P20 and P30 all correlate significantly with δ ln TX,
though with non-negligible scatter. In all cases this scat-
ter is larger for δ ln TSL than it is for δ ln TEW. Next, we
considered the possibility that M -TX scatter is driven by
low-mass clusters. We tested the degree to which scatter
can be reduced by filtering out these systems. This con-
sisted of performing a cut at 2 keV, for which we saw that
it yielded a modest improvement in mass estimates. To
see whether incorporating substructure could refine these
mass estimates, we first showed that w, η, P20, and P30
correlate significantly with the difference δ ln M between
masses predicted from the mean M (TX) relation and
the true cluster masses, with non-negligible scatter that
again is less for M (TEW) than it is for M (TSL). Then we
adopted a full three-parameter model, M -TX-S, which
includes substructure information S estimated using w,
η, P20, and P30. Scatter about the basic two-parameter
M -TEW relation was 0.094. Including substructure as a
third parameter reduced the scatter to 0.072 for centroid
variation, 0.084 for axial ratio, 0.081 for P20, and 0.084
for P30. Scatter about the basic two-parameter M -TSL
relation was 0.124, and including substructure as a third
parameter reduced the scatter to 0.085 for centroid vari-
ation, 0.112 for axial ratio, 0.110 for P20, and 0.108 for
P30. As one last test, and to increase our confidence
that our substructure measures are not relying on po-
tentially non-physical core structure in the simulations,
we also repeated the comparison of mass-estimates for
TEW, with the core regions of the clusters excised. First,
removing the core slightly increased the scatter in M -TX
possibly by making the average temperature more sen-
sitive to structure outside the core. Second, even with
the cores removed the improvement in mass-estimates
obtained using substructure information remains. Based
on these results, it appears that centroid variation is the
best substructure statistic to use when including a sub-
structure correction in the M -TEW relation. However,
the correlations we have found in this sample of simu-
lated clusters might not hold in samples of real clusters,
because relaxed clusters in the real universe tend to have
cooler cores than our simulated clusters do.
The authors wish to thank Stefano Borgani for con-
tributing the simulation data on which this project was
based and for his helpful comments on the manuscript.
This work was supported by NASA through grants
NNG04GI89G and NNG05GD82G, through Chandra
theory grant TM8-9010X, and through Chandra archive
grant SAOAR5-6016X.
REFERENCES
Borgani, S., Murante, G., Springel, V., Diaferio, A., Dolag, K.,
Moscardini, L., Tormen, G., Tornatore, L., & Tozzi, P. 2004,
MNRAS, 348, 1078
Buote, D. A., & Tsai, J. C. 1995, ApJ, 452, 522
Evrard, A. E., Bialek, J., Busha, M., White, M., Habib, S.,
Heitmann, K., Warren, M., Rasia, E., Tormen, G., Moscardini,
L., Power, C., Jenkins, A. R., Gao, L., Frenk, C. S., Springel,
V., White, S. D. M., & Diemand, J. 2007, ArXiv Astrophysics
e-prints
Evrard, A. E., Mohr, J. J., Fabricant, D. G., & Geller, M. J. 1993,
ApJ, 419, L9+
Fabian, A. C., Crawford, C. S., Edge, A. C., & Mushotzky, R. F.
1994, MNRAS, 267, 779
Hu, W. 2003, Phys. Rev. D, 67, 081304
Jeltema, T. E., Hallman, E. J., Burns, J. O., & Motl, P. M. 2007,
ArXiv e-prints, 708
Kaiser, N. 1986, MNRAS, 222, 323
Kay, S. T., da Silva, A. C., Aghanim, N., Blanchard, A., Liddle,
A. R., Puget, J.-L., Sadat, R., & Thomas, P. A. 2007, MNRAS,
377, 317
Kravtsov, A. V., Vikhlinin, A., & Nagai, D. 2006, ApJ, 650, 128
Levine, E. S., Schulz, A. E., & White, M. 2002, ApJ, 577, 569
10
Lima, M., & Hu, W. 2005, Phys. Rev. D, 72, 043006
Majumdar, S., & Mohr, J. J. 2004, ApJ, 613, 41
Markevitch, M. 1998, ApJ, 504, 27
Mathiesen, B. F., & Evrard, A. E. 2001, ApJ, 546, 100
Mazzotta, P., Rasia, E., Moscardini, L., & Tormen, G. 2004, ArXiv
Astrophysics e-prints
McCarthy, I. G., Balogh, M. L., Babul, A., Poole, G. B., & Horner,
D. J. 2004, ApJ, 613, 811
Nagai, D., Kravtsov, A. V., & Vikhlinin, A. 2007, ApJ, 668, 1
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1995, MNRAS,
275, 720
Riess, A. G., Press, W. H., & Kirshner, R. P. 1996, ApJ, 473, 88
Saro, A., Borgani, S., Tornatore, L., Dolag, K., Murante, G.,
Biviano, A., Calura, F., & Charlot, S. 2006, MNRAS, 373, 397
Springel, V. 2005, MNRAS, 364, 1105
Tormen, G. 1997, MNRAS, 290, 411
Tornatore, L., Borgani, S., Springel, V., Matteucci, F., Menci, N.,
& Murante, G. 2003, MNRAS, 342, 1025
Valdarnini, R. 2006, New Astronomy, 12, 71
Vikhlinin, A. 2006, ApJ, 640, 710
Voit, G. M. 2005, Phys. Rev. D, 77, 207
Voit, G. M., Bryan, G. L., Balogh, M. L., & Bower, R. G. 2002,
O'Hara, T. B., Mohr, J. J., Bialek, J. J., & Evrard, A. E. 2006,
ApJ, 576, 601
ApJ, 639, 64
Wang, L., & Steinhardt, P. J. 1998, ApJ, 508, 483
Pearce, F. R., Thomas, P. A., & Couchman, H. M. P. 1994,
MNRAS, 268, 953
Phillips, M. M. 1993, ApJ, 413, L105
Poole, G. B., Babul, A., McCarthy, I. G., Fardal, M. A., Bildfell,
C. J., Quinn, T., & Mahdavi, A. 2007, MNRAS, 380, 437
|
astro-ph/0503310 | 1 | 0503 | 2005-03-15T09:48:33 | Discovery of a Fifth Image of the Large Separation Gravitationally Lensed Quasar SDSS J1004+4112 | [
"astro-ph"
] | We report the discovery of a fifth image in the large separation lensed quasar system SDSS J1004+4112. A faint point source located 0.2'' from the center of the brightest galaxy in the lensing cluster is detected in images taken with the Advanced Camera for Surveys (ACS) and the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) on the Hubble Space Telescope. The flux ratio between the point source and the brightest lensed component in the ACS image is similar to that in the NICMOS image. The location and brightness of the point source are consistent with lens model predictions for a lensed image. We therefore conclude that the point source is likely to be a fifth image of the source quasar. In addition, the NICMOS image reveals the lensed host galaxy of the source quasar, which can strongly constrain the structure of the lensing critical curves and thereby the mass distribution of the lensing cluster. | astro-ph | astro-ph | PASJ: Publ. Astron. Soc. Japan , 1 -- ??,
c(cid:13) 2018. Astronomical Society of Japan.
5
0
0
2
r
a
M
5
1
1
v
0
1
3
3
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Discovery of a Fifth Image of the Large Separation Gravitationally
Lensed Quasar SDSS J1004+4112∗
Naohisa Inada,1 Masamune Oguri,2,3 Charles R. Keeton,4 Daniel J. Eisenstein,5
Francisco J. Castander,6 Kuenley Chiu,7 Joseph F. Hennawi,8 David E. Johnston,2
Bartosz Pindor,9 Gordon T. Richards,2 Hans-Walter Rix,10 Donald P. Schneider,11 Wei Zheng7
1Institute of Astronomy, Faculty of Science, University of Tokyo, 2-21-1 Osawa, Mitaka, Tokyo 181-0015.
2Princeton University Observatory, Peyton Hall, Princeton, NJ 08544, USA.
3Department of Physics, University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033.
4Department of Physics and Astronomy, Rutgers University, 136 Frelinghuysen Road, Piscataway, NJ 08854, USA.
5Steward Observatory, University of Arizona, 933 North Cherry Avenue, Tucson, AZ 85721, USA.
6Institut d'Estudis Espacials de Catalunya/CSIC, Gran Capita 2-4, 08034 Barcelona, Spain.
7Department of Physics and Astronomy, Johns Hopkins University,
3701 San Martin Drive, Baltimore, MD 21218, USA
8Department of Astronomy, University of California at Berkeley, 601 Campbell Hall, Berkeley, CA 94720, USA.
9Department of Astronomy, University of Tronto, 60 St. George Street, Tronto, Ontario M5S 3H8, Canada.
10Max-Planck Institute for Astronomy, Konigstuhl 17, D-69117 Heidelberg, Germany.
11Department of Astronomy and Astrophysics, The Pennsylvania State University,
525 Davey Laboratory, University Park, PA 16802, USA.
(Received 2004 October 27; accepted 2005 March 14)
Abstract
We report the discovery of a fifth image in the large separation lensed quasar system SDSS J1004+4112.
A faint point source located 0.′′2 from the center of the brightest galaxy in the lensing cluster is detected
in images taken with the Advanced Camera for Surveys (ACS) and the Near Infrared Camera and Multi-
Object Spectrometer (NICMOS) on the Hubble Space Telescope. The flux ratio between the point source
and the brightest lensed component in the ACS image is similar to that in the NICMOS image. The
location and brightness of the point source are consistent with lens model predictions for a lensed image.
We therefore conclude that the point source is likely to be a fifth image of the source quasar. In addition,
the NICMOS image reveals the lensed host galaxy of the source quasar, which can strongly constrain the
structure of the lensing critical curves and thereby the mass distribution of the lensing cluster.
Key words: galaxies: quasars:
individual (SDSS J1004+4112) -- gravitational lensing -- Galaxy:
structure
1.
Introduction
Gravitational lensing is a unique tool for exploring the
distribution of matter, particularly that of dark matter.
The recently discovered largest separation lensed quasar,
SDSS J1004+4112 (Inada et al. 2003; Oguri et al. 2004),
has opened a new window for probing dark matter dis-
tributions in the universe. The quasar was discovered in
the Sloan Digital Sky Survey (SDSS; York et al. 2000;
Abazajian et al. 2004), and the lensing hypothesis was
confirmed by subsequent observations with the Subaru
8.2-m telescope and the Keck telescope. The system con-
sists of four lensed quasar components (i′ =18.5, 18.9,
19.4, and 20.1) at z = 1.734 and the maximum separa-
tion angle between the lensed images is 14.′′62. The lens-
ing object must be a massive object, such as a cluster of
* Based on observations with the NASA/ESA Hubble Space
Telescope, obtained at the Space Telescope Science Institute,
which is operated by the Association of Universities for Research
in Astronomy, Inc., under NASA contract NAS 5-26555. These
observations are associated with HST program 9744.
galaxies, to produce such a large image separation; indeed,
we have identified a z = 0.68 cluster centered among the
four lensed images. The discovery of a single, cluster-size
lensed quasar among the current SDSS quasars is consis-
tent with the theoretical expectation of lensing based on
the cold dark matter model (Oguri et al. 2004; Oguri &
Keeton 2004).
SDSS J1004+4112 is unique in the sense that 1) the
quadruple images place robust constraints on the inner-
most region of the lensing cluster, and 2) the lensing clus-
ter is a strong lensing selected cluster of galaxies. These
features indicate that the mass modeling of this lens sys-
tem may offer valuable information on the structure of
clusters of galaxies. The first attempt at modeling the
system with various parametric models revealed the elon-
gated and complicated mass distribution of the lensing
cluster (Oguri et al. 2004). Williams & Saha (2004) re-
cently studied this system using a free-form reconstruc-
tion technique and reached similar conclusions. However,
important degeneracies between different models remain,
and further follow-up observations are required to deter-
2
INADA ET AL.
mine the mass distribution more precisely.
Observations using the Hubble Space Telescope (HST)
can offer such new data.
In particular, high-resolution
HST images could be quite effective at detecting lensed
images of the quasar host galaxy, arc or arclet images
of other lensed sources, and perhaps a central or "odd"
image of the lensed quasar system that would be ex-
pected theoretically (e.g., Burke 1981; Rusin 2002). A
central image is especially useful for providing tight con-
straints on the central mass distribution of the lensing
object. This was indeed demonstrated by the first discov-
ery of a central image in a lensed quasar system; Winn,
Rusin, & Kochanek (2004) showed that the central image
of PMN J1632−0033 requires β = 1.91 ± 0.02 (2σ confi-
dence) when a power-law density profile ρ(r) ∝ r−β is as-
sumed. Possible central images have also been identified in
lensed arc systems, such as CL 0024+1654 (Colley, Tyson,
& Turner 1996), MS 2137.3−2353 (Gavazzi et al. 2003),
and A1689 (Broadhurst et al. 2005), and they also provide
important constraints on mass models (see, e.g., Gavazzi
et al. 2003).
In this
letter, we present an identification of a
fifth image of the lensed quasar with the Advanced
Camera for Surveys (ACS; Clampin 2000) and the
Near Infrared Camera and Multi-Object Spectrometer
(NICMOS; Thompson 1992) installed on the HST. In ad-
dition, we report unambiguous detection of the lensed host
galaxy in the NICMOS image.
2. The ACS Observation
An ACS observation (with the Wide Field Channel) in
the F814W filter (≈I-band) was conducted on 2004 April
28, under the program "HST Imaging of Gravitational
Lenses" (GO-9744, PI C. Kochanek). The observation
consisted of five dithered exposures taken in ACCUM
mode. The total exposure time was 405 seconds. The
reduced (drizzled and calibrated) images were extracted
using the CALACS pipeline (Hack 1999), which includes
the PyDrizzle algorithm. We further rejected cosmic rays
using the L.A.Cosmic package (van Dokkum 2001) in the
drizzled image. A 40′′ × 40′′ subsection of a median of
the five dithered images is shown in Figure 1. Following
Figure 9 of Oguri et al. (2004), the four lensed components
are denoted as "A -- D", and three central bright galaxies
of the lensing cluster are denoted as "G1 -- G3". The red-
shifts of galaxies G1 -- G3 are 0.680, 0.675, and 0.675, re-
spectively (Oguri et al. 2004). The relative positions of
components A -- D were calculated by a single Gaussian fit,
and the flux ratios of components A -- D were estimated by
fitting the PSF stars produced by the Tiny Tim software
(version 6.1a; Krist & Hook 2003), in the drizzled (and
cosmic ray rejected) image. The PSF of a quasar was
constructed with an αν = −0.5 power law spectrum in the
F814W wavelength region (corresponding to ∼3000 A in
the rest frame). The relative positions derived from the
ACS (F814W) are consistent with those derived from the
Subaru i′ band image (Oguri et al. 2004) within ∼5σ. The
four components are all unsaturated, and the AB magni-
Table 1. Relative
SDSS J11004+4112 in the HST ACS Image
Positions
and
Flux Ratios
Object
y[arcsec]∗ Flux Ratio†
x[arcsec]∗
0.0000±0.001
-1.317±0.002
11.039±0.002
8.399±0.004
7.197±0.009
7.114±0.030
0.0000±0.001
3.532±0.002
-4.492±0.002
9.707±0.004
4.603±0.009
4.409±0.030
A
B
C
D
E
G1
[Vol. ,
of
1.000
0.732
0.346
0.207
0.003
--
∗ The positive directions of x and y are defined by West and North,
respectively.
† Errors of fitting by quasar PSFs are about 10%. The error of com-
ponent E might be much larger due to the over-subtraction of G1.
tude of component A in the F814W filter is estimated to
be 18.4. The position of component G1 was extracted by
the Source Extractor algorithm (Bertin & Arnouts 1996).
The results are summarized in Table 1.
3. The NICMOS Observation
The NICMOS imaging observation was also conducted
under the same HST program on 2004 October 9. The
observation consists of four dithered exposures taken in
MULTIACCUM mode, using the F160W filter (≈H-
band). The exposure time was 640 sec for two of the
exposures and 704 sec for the other two. The calibration
was extracted by the CALNICA pipeline, and the cen-
tral bad columns of each dithered image were corrected
by linear interpolation. The combined image is shown in
Figure 2. First, we confirm a galaxy near component A
(marked as G4 in Figure 2), which may host the star or
stars responsible for microlensing of the broad emission
line region (Richards et al. 2004). In addition, extended
emission is clearly seen around components A, B and C.
Although such extensions are also seen in the ACS im-
ages (see Figure 1), their existence is much more robust
in the NICMOS image. The fact that these extensions are
obvious in the F160W (near-infrared) image and faint in
the F814W (optical) image, and that the distortions agree
with the theoretical critical curves (see Figure 17 of Oguri
et al. 2004), demonstrates that the extended flux is due to
the lensed host galaxy of the source quasar. These images
provide many new constraints on lens models that will sig-
nificantly improve our ability to determine the mass dis-
tribution of the lensing cluster. They do make lens mod-
eling more computationally intensive (because one must
account for the intrinsic shape of the host galaxy, and also
for the effects of the point spread function), so we defer
detailed modeling of the arcs to a subsequent paper.
4. Fifth Image
Of interest is the existence of a point source near the
center of G1. The left panel in Figure 3 shows the ACS im-
age of galaxy G1; what appears to be an unresolved source
is clearly seen approximately 0.′′2 northwest of the center
of G1. This feature is neither a bad pixel nor a cosmic ray;
No. ]
Fifth image of SDSS J1004+4112
3
Fig. 1. 40′′ ×40′′ subsection of the median combined ACS image of SDSS J1004+4112. The four quasar images are labeled A -- D. The
galaxies of the lensing cluster labeled G1 -- G3 have redshifts of 0.680, 0.675, and 0.675, respectively. The pixel scale is approximately
0.′′05 pixel−1.
the source is seen in all dithered images. The right panel
in Figure 3 displays the image after subtracting the signal
from G1 (modeled with the GALFIT package of Peng et
al. 2002). Due to the existence of the unresolved source
near the center of G1 (the peak flux of this unresolved
source is almost same as that of G1), G1 was slightly
over-subtracted. However, we can see a new source, la-
beled E, in the subtracted image. This object is classified
as a point source by the Source Extractor algorithm. The
position and brightness of E, based on a single Gaussian
fit to the data, are given in Table 1.
We find component E also in the NICMOS image; the
left panel in Figure 4 is the NICMOS image of galaxy
G1. We subtracted the signal from G1 using the GALFIT
package, which is shown in the right panel of Figure 4.
We confirm component E in the subtracted image, al-
though G1 was slightly over-subtracted as in the ACS
image. Measuring the flux of component E with a sin-
gle Gaussian fit, we find that the flux ratios between E
and A (E/A) in the ACS and NICMOS images are 0.003
and 0.004, respectively. This remarkable agreement of the
flux ratios supports the idea that the point source is a fifth
image of the lensed quasar.
To test this hypothesis, we have refined the lens models
presented in Oguri et al. (2004) using the more precise
HST data. The models consist of a singular isothermal
ellipsoid mass distribution for galaxy G1, and an NFW
(Navarro, Frenk, & White 1997) elliptical potential for the
cluster. We demand that the models reproduce the rela-
tive positions and brightnesses of components A -- D (and
the relative position of G1), as well as the position of
component E; we do not use the brightness of E as a con-
straint, because we want to see what the models predict.
Adopting the same approach as Oguri et al. (2004), we
use the lensmodel software (Keeton 2001) for Monte Carlo
sampling of the parameter space. There is a wide range
of models consistent with the data, indicating that it is
not difficult to produce a 5-image lens matching the con-
figuration of components A -- E, and that significant model
degeneracies remain. More interesting are the model pre-
dictions for the flux ratio between the fifth image and
component A, shown in Figure 5. The predictions span
a remarkable nine orders of magnitude, from E/A ∼ 0.1
down to E/A ∼ 10−10, but a significant fraction predict
E/A in the range 0.001 -- 0.01, consistent with the observed
value. In other words, there are (many) reasonable mod-
els that can fit all of the HST data under the hypothesis
that E is a fifth image; conversely, the observed properties
of E are highly compatible with that hypothesis.
Given the enormous range of model predictions for the
brightness of E, it appears that the observed brightness
offers strong constraints on the models. We caution that
one must be careful in using the brightness of E as a con-
straint, because its measured flux could be contaminated
by improper subtraction of the galaxy or by physical ef-
fects such as microlensing or extinction. Nevertheless, the
range of predictions is so large that even conservative es-
timates of systematic uncertainties should still yield very
4
INADA ET AL.
[Vol. ,
Fig. 4. Left: The NICMOS image of central galaxy G1.
Right: The NICMOS image after subtracting galaxy G1. The
stellar object near the center of G1 can be also seen in the
NICMOS image.
lower bound on β, at least over the range 1.6 ≤ β ≤ 2.0
that we have explored so far. Apparently the complexity
of the SDSS J1004+4112 lens potential, with a cluster in
addition to the galaxy, prevents a unique measurement
of the value of β. Nevertheless, it is clear that compo-
nent E provides important new constraints on the mass
distribution of this interesting lens system.
5. Summary
We have presented the HST ACS and NICMOS images
of SDSS J1004+4112, which reveal a fifth image of the
lensed quasar core. The fifth image offers a unique probe
of the mass distribution of the cluster core. Deep spectro-
scopic observations of component E with large telescopes,
such as the Subaru Telescope, offer the best prospect for
the final confirmation that it is a lensed quasar image. In
the NICMOS image, we also found unambiguous evidence
of the lensed host galaxy of the source quasar. These
extended images provide strong additional constraints on
mass models of the lensing cluster, which are expected to
break degeneracies seen in the modeling studies to date. A
detailed analysis of lens models including the host galaxy
images is underway and will be presented elsewhere.
A portion of this work was supported by NASA HST-
GO-09744.20. NI and MO are supported by JSPS through
JSPS Research Fellowship for Young Scientists.
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 Institutions. The
Participating Institutions re The University of Chicago,
Fermilab, the Institute for Advanced Study, the Japan
Participation Group, The Johns Hopkins University, the
Korean Scientist Group, Los Alamos National Laboratory,
the Max-Planck-Institute for Astronomy (MPIA), the
Max-Planck-Institute
(MPA), New
for Astrophysics
Fig. 2. The combined NICMOS image of SDSS J1004+4112.
The pixel scale is approximately 0.′′075 pixel−1. The lensed
host galaxy is seen more prominently in this NICMOS image
than in the ACS image.
In addition, we can see G4 near
component A.
Fig. 3. Left: The ACS image of central galaxy G1. Right:
The ACS image after subtracting galaxy G1. The residual
image clearly shows a stellar object near the center of G1.
interesting results. As an example, to test the ability
of E to constrain the central density profile of G1, we
switch from an isothermal model to a more general power
law density profile ρ(r) ∝ r−β for the galaxy (for com-
putational simplicity, we now assume that the potential,
rather than the density, has elliptical symmetry). We find
that there is a wide range of models with 1.6 ≤ β ≤ 2.0
that predict 0.001 <
∼ 0.01. By contrast, all of the
models we examined with β ≥ 2.1 predict E/A <
∼ 10−10,
which grossly contradicts even a conservative reading of
the data. In other words, the observed properties of E im-
ply that the galaxy mass distribution cannot be steeper
than isothermal (i.e., β ≤ 2). This upper bound is similar
to that found by Winn, Rusin, & Kochanek (2004) from
the central image in PMN J1632−0033 (specifically, they
found β = 1.91 ± 0.02). However, SDSS J1004+4112 dif-
fers from PMN J1632−0033 in that we do not obtain any
∼ E/A <
No. ]
Fifth image of SDSS J1004+4112
5
observed value in ACS
0.2
Rusin, D. 2002, ApJ, 572, 705
Thompson, R. 1992, Space Science Reviews (ISSN 0038-6308),
61, no. 1-2, 69
Williams, L. L. R. & Saha, P. 2004, AJ, 128, 2631
Winn, J. N., Rusin, D., & Kochanek, C. S. 2004, Nature, 427,
613
York, D. G., et al. 2000, AJ, 120, 1579
l
s
e
d
o
m
f
o
n
o
i
t
c
a
r
f
0.1
0
-- 10
-- 5
0
predicted log10(E/A)
Fig. 5. Frequency (fraction of models) of the predicted flux
ratio of E/A from two component (central galaxy + cluster
component) lens models. The models are constrained by the
observed positions and brightnesses of components A -- D (and
the observed position of G1), and the position of component
E. We assume that the scale radius of the NFW component is
40.′′0, but the results are insensitive to the particular value.
The gray filled circle represents the observed value of E/A (in
ACS) with 50% error.
Mexico State University, University of Pittsburgh,
University of Portsmouth, Princeton University,
the
United States Naval Observatory, and the University of
Washington.
References
Abazajian, K. 2004, AJ, 128, 502
Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
Broadhurst, T., et al. 2005, ApJ, in press (astro-ph/0409132)
Burke, W. L. 1981, ApJL, 244, L1
Clampin, M. et al. 2000, SPIE, 401, 344
Colley, W. N., Tyson, J. A., & Turner, E. L. 1996, ApJL, 461,
L83
van Dokkum P. G. 2001, PASP, 113, 1420
Gavazzi, R., Fort, B., Mellier, Y., Pello, R., & Dantel-Fort, M.
2003, A&A, 403, 11
Hack, W. J. 1999, CALACS Operation and Implementation,
ISR ACS-99-03
Inada, N., et al. 2003, Nature, 426, 810
Keeton, C. R. 2001, preprint (astro-ph/0102340)
Krist, J. E. & Hook, R. N. 2003, The Tiny Tim User's Guide,
Version 6.1a
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ,
490, 493
Oguri, M., et al. 2004, ApJ, 605, 78
Oguri, M., & Keeton, C. R. 2004, ApJ, 610, 663
Peng, C. Y., Ho, L. C., Impey, C. D., & Rix, H.-W. 2002, AJ,
124, 266
Richards, G. T., et al. 2004, ApJ, 610, 679
|
astro-ph/0410031 | 2 | 0410 | 2005-03-07T21:14:29 | The Cosmology of Generalized Modified Gravity Models | [
"astro-ph",
"hep-ph",
"hep-th"
] | We consider general curvature-invariant modifications of the Einstein-Hilbert action that become important only in regions of extremely low space-time curvature. We investigate the far future evolution of the universe in such models, examining the possibilities for cosmic acceleration and other ultimate destinies. The models generically possess de Sitter space as an unstable solution and exhibit an interesting set of attractor solutions which, in some cases, provide alternatives to dark energy models. | astro-ph | astro-ph | The Cosmology of Generalized Modified Gravity Models
Sean M. Carroll1 ∗ , Antonio De Felice2 † , Vikram Duvvuri1 ‡ ,
Damien A. Easson2 § , Mark Trodden2 ¶ and Michael S. Turner1,3,4 ∗∗
1Enrico Fermi Institute, Department of Physics,
and Kavli Institute for Cosmological Physics, University of Chicago,
5640 S. El lis Avenue, Chicago, IL 60637-1433, USA.
2Department of Physics, Syracuse University, Syracuse, NY 13244-1130, USA.
3Department of Astronomy & Astrophysics,
University of Chicago, Chicago, IL 60637-1433, USA.
4NASA/Fermilab Astrophysics Center,
Fermi National Accelerator Laboratory, Batavia, IL 60510-0500, USA.
Abstract
We consider general curvature-invariant modifications of the Einstein-Hilbert action that become
important only in regions of extremely low space-time curvature. We investigate the far future
evolution of the universe in such models, examining the possibilities for cosmic acceleration and
other ultimate destinies. The models generically possess de Sitter space as an unstable solution
and exhibit an interesting set of attractor solutions which, in some cases, provide alternatives to
dark energy models.
5
0
0
2
r
a
M
7
2
v
1
3
0
0
1
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
∗ [email protected]
† [email protected]
‡ [email protected]
§ [email protected]
¶ [email protected]
∗∗ [email protected]
1
I.
INTRODUCTION
The acceleration of the universe presents one of the greatest problems in theoreti-
cal physics today. The increasingly accurate observations of type Ia supernovae light-
curves, coupled with exquisite measurements of CMB anisotropies and large scale structure
data [1, 2, 3, 4, 5, 6] have forced this issue to the forefront of those facing particle physicists,
cosmologists and gravitational physicists alike.
This problem has been attacked head on, but no compelling, well-developed and well-
motivated solutions have yet emerged. While much work has focused on the search for new
matter sources that yield accelerating solutions to general relativity, more recently some
authors have turned to the complementary approach of examining whether new gravitational
physics might be responsible for cosmic acceleration.
There have been a number of different attempts [7, 8, 9, 10, 11, 12, 13, 14, 15] to modify
gravity to yield accelerating cosmologies at late times. The path we are concerned with
in this paper is the direct addition of higher order curvature invariants to the Einstein-
Hilbert action. The first example of this was provided by the model of Carroll, Duvvuri,
Trodden, and Turner (CDTT) [7] (see also [8]). For subsequent work on various aspects and
extensions of this model, see [16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32].
In particular, the simplest model has been shown to conflict with solar system tests of
gravity [19, 22, 25]. Our approach is purely phenomenological. The evidence for cosmic
acceleration is very sound. In pure Einstein gravity, as matter dilutes away in the expanding
universe, the expansion rate inevitably slows. Our question is can this be avoided within the
gravitational sector of the theory, as opposed to adding new energy sources. One way is by
adding a cosmological constant. In this paper we explore a wider class of modifications that
share the feature of late-time accelerating behavior, thus fitting cosmological observations.
Consider a simple correction to the Einstein-Hilbert action,
µ4
M 2
R (cid:19) + Z d4x √−g LM .
2 Z d4x √−g (cid:18)R −
P
where µ is a new parameter with units of [mass] and LM is the Lagrangian density for matter.
This action gives rise to fourth-order equations of motion. In the case of an action depending
S =
(1)
2
purely on the Ricci scalar (and on its derivatives) it is possible to transform from the frame
used in (1), which we call the matter frame, to an Einstein frame, in which the gravitational
Lagrangian takes the Einstein-Hilbert form and the additional degrees of freedom ( H and
H ) are represented by a scalar field φ. The details of this can be found in [7]. The scalar
field is minimally coupled to Einstein gravity, non-minimally coupled to matter, and has a
potential given by
(2)
MP ! vuutexp r 2
MP ! − 1 .
P exp −2r 2
φ
φ
V (φ) = µ2M 2
3
3
Consider vacuum cosmological solutions. We must specify the initial values of φ and φ′ ,
i . For simplicity we take φi ≪ MP . There are three qualitatively distinct
outcomes, depending on the value of φ′
i .
1. Eternal de Sitter. There is a critical value of φ′
i ≡ φ′
maximum of the potential V (φ) and comes to rest. In this case the Universe asymptotically
C for which φ just reaches the
denoted as φi and φ′
evolves to a de Sitter solution. This solution requires tuning and is unstable, since any
perturbation will induce the field to roll away from the maximum of its potential.
2. Power-Law Acceleration. For φ′
C , the field overshoots the maximum of V (φ)
i > φ′
and the Universe evolves to late-time power-law inflation, with observational consequences
similar to dark energy with equation-of-state parameter wDE = −2/3.
3. Future Singularity. For φ′
i < φ′
C , φ does not reach the maximum of its potential and
rolls back down to φ = 0. This yields a future curvature singularity.
In the more interesting case in which the Universe contains matter, it is possible to show
that the three possible cosmic futures identified in the vacuum case remain in the presence
of matter.
By choosing µ ∼ 10−33 eV, the corrections to the standard cosmology only become im-
portant at the present epoch, making this theory a candidate to explain the observed accel-
eration of the Universe without recourse to dark energy. Since we have no particular reason
for choosing this value of µ, such a tuning is certainly not attractive. However, it is worth
commenting that this small correction to the action, unlike most small corrections in physics,
is destined to be important as the Universe evolves.
3
Clearly our choice of correction to the gravitational action can be generalized. Terms of
the form −µ2(n+1) /Rn , with n > 1, lead to similar late-time self acceleration, with behavior
similar to a dark energy component with equation of state parameter
2(n + 2)
3(2n + 1)(n + 1)
weff = −1 +
Therefore, such modifications can easily accommodate current observational bounds [33, 34]
on the equation of state parameter −1.45 < wDE < −0.74 (95% confidence level). In the
asymptotic regime, n = 1 is ruled out at this level, while n ≥ 2 is allowed; even n = 1 is
permitted if we are near the top of the potential.
(3)
.
In this paper we seek to extend this approach. For the actions considered in this paper,
a tranformation between the matter and Einstein frames does not necessarily make sense.
Therefore, we would like to analyze the dynamics in the matter frame itself. A technique for
this analysis is presented here.
As an example, we begin by applying this technique to the CDTT model. We will work
in the matter frame with a spatially flat Robertson-Walker metric
ds2 = −dt2 + a(t)2 d~x2 ,
with a(t) being the scale factor. The time-time component of the field equations, the Fried-
(4)
mann equation in this non-standard cosmology, is
µ4
12( H + 2H 2)3 (cid:16)2H H + 15H 2 H + 2 H 2 + 6H 4(cid:17) =
3H 2 −
where an overdot denotes differentiation with respect to cosmic time, and H = a/a.
To perform a phase-space analysis of such equations we write H as a function of H . Let
ρM
M 2
P
(5)
,
x = −H (t)
y = H (t) ,
(6)
so that
d H
d H
dy
dH
= −y
dx
dt
dH
dt
In this way we may write (5), a third order equation in a(t), as a second order equation
H =
(7)
=
.
in H (t) and hence as a first order equation in y (x). The fact that the FRW equation for
4
any f (R) theory can be reduced to the first-order equation for the spatially flat case (second
order in case of a non-zero spatial curvature) was first introduced in [35], [36] .
Many cosmologically interesting solutions, including accelerating ones, are power-law so-
lutions of the form a(t) ∝ tp . In such cases H = −H 2/p (i.e. y = −x2/p). In anticipation
of finding such solutions as asymptotic solutions to our equations, we define a new function
v (x) by
x2
v (x) = −
y
with v 6= 0. Power-law solutions in the asymptotic future are then easily identified as
v (x) → p = constant as H → 0.
Furthermore, if
(8)
,
H = y → ∞,
if x 6= 0.
v → 0
So if x is not zero, as v → 0 we approach the singularity H → ∞. This trick is invoked
throughout this paper.
then
(9)
x
dv
dx
Here, let us apply it to the simple case of (5). The relevant first order equation is
1
2µ4 (cid:2)x4 (36 − 216v + 432v 2 − 288v 3) + µ4 (2v − 15v 2 + 6v 3)(cid:3) ,
with the resulting phase plot shown in figure 1.
Since the x-axis on this plot is (minus) the Hubble parameter, earlier times in the universe
= 2v +
(10)
lie to the (negative) left and late times lie closer to x = 0 (in all but exponential or phantom
evolution).
Note that the numerical solution shows the accelerating attractor a(t) ∝ t2 , corresponding
to v (x) → 2 as x → 0, as expected from the analytic, Einstein-frame method. Indeed, for
general n this attractor, at v = (2n + 1)(n + 1)/(n + 2) can be obtained directly from the
asymptotic form of the generalization of equation (10). In addition, our method pinpoints a
singularity in the phase space, corresponding to a power-law evolution with exponent p = 1/2
(that of radiation). Both features are also evident from the study of the asymptotically late
time behavior of equation (10). In order to have a constant v0 as a solution, we need
2v 2
0 − 5v0 + 2 = 0 ,
5
(11)
4
3
2
v
1
-0.8
-0.6
-0.4
-0.2
H
dHdt
0.8
0.6
0.4
0.2
-0.4
-0.3
-0.2
x
-0.1
0
0
FIG. 1: Two phase portraits for the modified gravity model proposed in [7]. The left portrait is in
the coordinates (x, v), for which an attractor at constant v = p corresponds to a power-law solution
with a(t) ∝ tp . The right portrait is for the same theory in the ( H , H ) plane, with the unstable de
Sitter solution at (0, 1).
which has two real solutions: v0 = 1/2, 2, as expected.
The singularity at p = 1/2 occurs because the Ricci scalar vanishes for this particular
power. One might worry that this is problematic for describing the radiation-dominated
phase in standard cosmology, since nucleosynthesis occurs during this epoch and provides a
particularly strong constraint on deviations from the standard Friedmann equation at that
time. However, as we shall see later, this is not a problem when matter sources are included
explicitly, since even during radiation domination the Ricci scalar does not vanish exactly,
but rather, has a small contribution from non-relativistic matter.
Nevertheless, the singularity is a new feature that was not found in the Einstein-frame
analysis [7]. This is presumably because R = 0 is a singular point of the conformal transfor-
mation used to reach the Einstein frame. We shall see similar singularities for some of the
more general actions we consider in this paper.
Another way to visualize the solutions to our models is to use a more traditional phase
6
portrait in the ( H , H ) plane.
The outline of this paper is as follows. In the next section we shall introduce the general
class of actions we are interested in. In section III we analyze the vacuum equations, describ-
ing the singularity and attractor structure in detail before moving on to some simple special
cases. In section IV we introduce matter into the equations, demonstrating briefly that the
late-time attractor solutions of the system remain unchanged and setting up the formalism
used in the appendix to establish stability of the system. In section V we summarize our
findings and comment on the status of these models as origins of cosmic acceleration. The
paper contains two appendices. Appendix A contains definitions of a number of functions
used in the body of the paper and Appendix B consists of a proof of the stability of the
vacuum solutions under the addition of matter.
II. A GENERAL NEW GRAVITATIONAL ACTION
We now generalize the action of [7] to include other curvature invariants. There are,
of course, any number of terms that we could consider. We have chosen to consider those
invariants of lowest mass dimension that are also parity-conserving
P ≡ Rµν Rµν
Q ≡ Rαβγ δ Rαβγ δ .
Since we are interested in adding terms to the action that explicitly forbid flat space as a
(12)
solution, we will, in a similar way as in [7], consider inverse powers of the above invariants.
It is likely that such terms introduce ghost degrees of freedom. We shall not address this
problem here, since it is beyond the scope of this paper. Rather, if ghosts arise we shall
require that some as yet unknown mechanism (for example, extra-dimensional effects) cut
off the theory in such a way that the associated instabilities do not appear on cosmological
time scales [37] (see [14] for an example of a concrete model where ghosts are brought under
control by higher-derivative terms). We therefore consider actions of the form
S = Z d4x√−g [R + f (R, P , Q)] + Z d4x √−g LM ,
7
(13)
where f (R, P , Q) is a general function describing deviations from general relativity.
It is convenient to define
fP ≡
fR ≡
in terms of which the equations of motion are
,
∂f
∂R
∂f
∂P
,
fQ ≡
∂f
∂Q
,
(14)
Rµν − 1
2 gµν R − 1
2 gµν f
+ fR Rµν + 2fP Rα
µ Rαν + 2fQ Rαβγµ Rαβγ
+ gµν (cid:3)fR − ∇µ∇ν fR − 2∇α∇β [fP Rα
+ gµν ∇α∇β (fP Rαβ ) − 4∇α∇β [fQ Rα
(µν )
It is straightforward to check that these equations reduce to those of the simple model
of [7] for f (R) = −µ4/R.
We would like to obtain constant curvature vacuum solutions to these field equations.
β ] = 8πG Tµν .
ν ) ] + (cid:3)(fP Rµν )
(µ δβ
(15)
ν
To do so, we take the trace of (15) and substitute Q = R2/4 and P = R2/6 (which are
identities satisfied by constant curvature spacetimes) into the resulting equation to obtain
the algebraic equation:
(2fQ + 3fP ) R2 + 6 (fR − 1) R − 12f = 0 .
Solving this equation for the Ricci scalar yields the constant curvature vacuum solutions.
(16)
Evidently, actions of the form (13) generically admit a maximally-symmetric solution:
R = a non-zero constant. However, an equally generic feature of such models is that this
deSitter solution is unstable. In the CDTT model the instability is to an accelerating power-
law attractor. This is a possibility that we will also see in many of the more general models
under consideration here.
Before we leave this section, note that, in analyzing the cosmology of these general models,
it is useful to have at hand the following expressions, which hold in a flat FRW background
R = 6 (cid:18) a2
a (cid:19) = 6 ( H + 2H 2)
a
a2 +
a2
a2
P = 12 (cid:18) a4
a (cid:19) = 12 [( H + H 2)2 + H 4 + H 2 ( H + H 2)]
a
a2 +
a4 +
a2
Q = 12 (cid:18) a4
a2
a2 (cid:19) = 12 [( H + H 2)2 + H 4 ] .
a4 +
8
(17)
(19)
(18)
We have provided these both in terms of the scale factor a(t) and in terms of the Hubble
parameter H (t) = a(t)/a(t), since they will be separately important in this paper.
III. VACUUM SOLUTIONS
In this section we study cosmological solutions to the field equations (15) in the absence of
matter sources. Physically, this is important because we are hoping to find novel cosmological
consequences arising purely from the gravitational sector of the theory. Mathematically, this
provides us with valuable insight into the structure of the equations, which take a significantly
simplified form wherein the Hubble parameter may be treated as the independent variable.
As mentioned in the previous section, we will consider inverse powers of our curvature
invariants and, for simplicity, we will specialize to a class of actions with
f (R, P , Q) = −
µ4n+2
(aR2 + bP + cQ)n ,
(20)
where n is a positive integer, µ has dimensions of mass and a, b and c are dimensionless
constants. [1] We will focus on the case n = 1 for most of the paper, because the analysis
is less involved for that case. For general n the qualitative features of the system are as for
n = 1 and we discuss the quantitative differences in our conclusions.
A. Distinguished Points of the Action (and Equations)
In general, the analogue of the Friedmann equation may be written in the following
convenient form
A + B H
C
where A = A(H, H ), B = B (H, H ) and C = C (H, H ) arise from the gravitational part of
= M ,
(21)
the action and M = M (a) describes the possible inclusion of matter.
[1] Another potentially interesting possibility is a correction of form f (R, P, Q) = µ4R
P . This term has the
same mass dimension as the µ4/R term of CDTT. However, it turns out that this model does not possess
any accelerating attractors. Specifically, the scale factor asymptotically approaches a(t) ∝ t0.37 .
9
It is also convenient to write this schematically in terms of our variables of the previous
section as
x
dv
dx
=
x6 f (v ) + µ6 v 2 g (v )
2µ6 v h(v )
,
(22)
where f (v ), g (v ) and h(v ) are 6th, 4th and 2nd degree polynomials respectively in the variable
v , whose explicit form is given in Appendix A. We are often interested in a particular subset
of the phase space, the region where v > 1/2. This is because from nucleosynthesis to the
matter-dominated epoch, we expect Einstein’s equations to provide a good approximation
to the dynamics, and therefore, when matter becomes subdominant we should have 1/2 <
v < 2/3.
There are three types of special points in the phase space plots of these equations:
1. Singular Points of the Friedmann Equation
In our introduction we reconsidered the model of [7] and discovered a singular point of
both the action and the equations of motion, corresponding to a power-law evolution
with exponent p = 1/2. In this particular case the singularity occurred because R ≡ 0
for p = 1/2. In our more general models, analogous singularities occur whenever the
denominator of the Friedmann equation blows up; i.e. at zeros of C , where C is defined
by (21). For the flat cosmological ansatz this occurs when
H !2
4 H
α
H + H 2 = −
,
where we have defined
(23)
α ≡
Recall that a, b, and c are parameters in the Lagrangian (20). In our variables of the
(24)
12a + 4b + 4c
12a + 3b + 2c
.
previous section this becomes
v = p1,2 ≡
each zero having multiplicity 3.
2 (cid:0)1 ± √1 − α(cid:1) ,
1
(25)
10
These singularities only exist if p1,2 are real, i.e. α ≤ 1 and therefore there are many
invariants that do not admit this type of singularity. In the simple case of [7], corre-
sponding to b = c = 0, note that we have α = 1 and recover pc = 1/2 as expected. If
α ≤ 1
p2 ≥ 1/2
p1 ≤ 1/2
2. Singular Points at which H → ∞
Points at which H → ∞ occur at zeros of B and, in our variables of the previous
section, correspond to
dx (cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
dv
→ +∞
at finite x and v . We denote the zeros by v (x) = v1 , v2 , where the vi are (in general
complex) constants constructed from a, b and c.
If g (v ) 6= 0 one singular point is at x = 0. Otherwise the singularities occur at solutions
of v h(v ) = 0, i.e.
(26)
(27)
v (cid:2)(108a2 + 51ab + 7bc + 30ac + 2c2 + 6b2 ) v 2
− (63ab + 6c2 + 9b2 + 108a2 + 15bc + 54ac) v
+18ab + 3c2 + 6bc + 27a2 + 18ac + 3b2 (cid:3) = 0 ,
which are given by
v1,2 =
2(4 − α) 1 ± r α − 1
3 !
3α
v3 = 0 .
(28)
(29)
3. Late time Stable Points.
Finally, we look for late time power-law attractors. This happens if v (x) → constant
(distinct from our previously mentioned singular values). Taking an asymptotic limit
11
of the equations of motion yields
v (cid:2)(288a2 + 18b2 + 8c2 + 144ab + 96ac + 24bc) v 4
− (1512a2 + 90b2 + 36c2 + 738ab + 468ac + 114bc) v 3
+(1836a2 + 123b2 + 62c2 + 951ab + 678ac + 175bc) v 2
−(846a2 + 69b2 + 44c2 + 489ab + 414ac + 113bc) v
+135a2 + 15b2 + 15c2 + 90ab + 90ac + 30bc(cid:3) = 0 , (30)
with solutions
(20 − 3α)2 !
120α
s1,2 =
20 − 3α
8
1 ± s1 −
s3 = p1
s4 = p2
s5 = 0 ,
(31)
(32)
(33)
and s2 > s1 .
It is clear that only s1 and s2 can be late-time power-law solutions. The other ones
represent singularities; v = 0 implies that y = H → ∞, whereas s3 and s4 are the
singular points we discussed earlier.
B. Summary of Possibilities
Here we summarize the different vacuum possibilities. It is useful to define the following
two constants
280 + 60√3
111 + 60√3 ≈ 1.78
η1 =
280 − 60√3
111 − 60√3 ≈ 24.9 .
1. α < 1. In this case vi are complex, whereas pi and si are real. It is straightforward to
η2 =
(34)
(35)
show that s2 > p2 > 1/2 and s2 > 1.
12
• 0 < α < 1. In this case 1/2 > s1 > p1 > 0 and 1 > p2 > 1/2. Solutions close to
p2 are repelled from it, whereas s1 is an attractor. This leads to decelerating late
time behavior.
• 0 > α > −160/9. Here s1 < p1 < 0 and solutions are attracted to (x = 0, v = 0).
Furthermore p2 > 1.
• α < −160/9.
attracted to (x = 0, v = 0). Again, p2 > 1.
In this case p1 < s1 < 0 and again we have that solutions are
2. α > 1. In this case vi are real, whereas pi are complex.
• 1 < α < 4/3. Here v1 < 1/2 and 1/2 < s1 < v2 < 1. Furthermore s2 > 1. Both
s1 and s2 are attractors, but s1 describes a decelerating phase.
• 4/3 < α < η1 . Here 1/2 < v1 < 1 and v2 > 1. Both (x = 0, v = 0) and s1,2 are
attractors, with the separatrix being at v1,2 .
• η1 < α < 4. In this case 1/2 < v1 < 1 and v2 > 1. Again, both (x = 0, v = 0)
and v2 are attractors and the separatrix is at v1 . There are no real solutions for
s1,2 .
• 4 < α < η2 . Here there are no real late-time attractors since the si are complex.
We also have v2 < 0 and v1 > 1. Solutions either evolve to (x = 0, v = 0) or to
v = +∞, with separatrix at v1 .
• α > η2 . In this case v2 < 0, v1 > 1 and s1 < s2 < 0. Evolution is to a decelerating
attractor.
3. α = 1. This yields a promising class of solutions. We have
1
2
All singularities occur as v (x) → 1/2. However, from nucleosynthesis onwards we
never encounter this point. It is simple to show that solutions evolve to a late-time
p1 = p2 = v1 = v2 = s1 =
.
(36)
power-law attractor describing an accelerating phase, with
s2 =
15
4
= 3.75 ,
13
(37)
which is otherwise independent of a, b, c.
The results of this section may be summarized in figure 2, showing the values of the
various distinguished points as α is varied.
5
4
3
2
1
0
0
s2
v2
v1
s1
1.5
2
2.5
p2
0.5
p1
1
α
FIG. 2: The values of the various distinguished points as α is varied.
C.
Inverse Powers of P ≡ Rµν Rµν
Let us begin by dealing only with actions containing modifications involving P ≡ Rµν Rµν .
Our prototype is to consider f (P ) = −m6/P , with m a parameter with dimensions of mass.
Using (16) we can see that there is a constant curvature vacuum solution to this action
given by
R(P )
const = (16)1/3 m2 .
(38)
However, we would like to investigate other cosmological solutions and analyze their stability.
From (15), with the flat cosmological ansatz, the analogue of the Friedmann equation
becomes
3H 2 −
m6
8(3H 4 + 3H 2 H + H 2)3 h H 4 + 11H 2 H 3 + 2H H 2 H
+ 33H 4 H 2 + 30H 6 H + 6H 3 H H + 6H 8 + 4H 5 H i = 0 .
14
(39)
We analyze this equation using the same technique, with the same definitions, as in the
example of the previous section. The relevant equation is
x
dv
dx
1
2m6v (2v 2 − 3v + 1) (cid:2)−x6 (24 − 216v + 864v 2 − 1944v 3
= 2v 2 +
+ 2592v 4 − 1944v 5 + 648v 6) + m6 (v 2 − 11v 3 + 33v 4 − 30v 5 + 6v 6)(cid:3) . (40)
The solution to this equation is displayed graphically in figure 3. We identify four fixed points
of the system; two attractors at v ≃ 0.77 and v ≃ 3.22 and two repellers at v ≃ 0.5 and
v = 1. Clearly, in order to obtain a late-time accelerating solution (p > 1), it is necessary to
give accelerating initial conditions (a > 0), otherwise the system is in the basin of attraction
of the non-accelerating attractor at p ≃ 0.77.
4
3
2
v
1
0
0
-0.01
-0.008
-0.006
-0.004
-0.002
x
-0.8
-0.6
-0.4
-0.2
H
1
0.8
0.6
0.4
0.2
dHdt
FIG. 3: Two phase portraits for the f (R, P , Q) = −m6/P modification. The left portrait is in the
coordinates (x, v), for which an attractor at constant v = p corresponds to a power-law solution
with a(t) ∝ tp . The right portrait is for the same theory in the ( H , H ) plane. There are two
late-time power law attractors corresponding to p = (4 ± √6)/2. The (-) branch, represented by
the solid lines, is non-accelerating, while the (+) branch, represented by the dash-dotted line, is
accelerating.
15
The exact exponents of the two late-time attractors of the system are obtained by studying
the asymptotic behavior of (40). Substituting in a power-law ansatz and taking the late-time
limit we find that, in order to have a constant v = v0 as a solution, the exponent must satisfy
6v 4
0 − 30v 3
0 + 41v 2
0 − 23v0 + 5 = 0 .
This equation has two real solutions (and two complex ones). The real solutions are: v0 =
2 − √6/2 ≃ 0.77 and v0 = 2 + √6/2 ≃ 3.22.
Even the non-accelerating attractor is of some interest in this model. In order for structure
(41)
to form in the universe, there must be a sufficiently long epoch of matter domination, for
which a(t) ∝ t2/3 . As matter redshifts away, however, since the universe is decelerating, we
expect the universe to approach the attractor at p ≃ 0.77. This corresponds to an effective
equation of state weff ≃ −0.13; i.e. negative pressure, although not negative enough to
provide a good fit to the supernova observations.
D.
Inverse Powers of Q ≡ Rαβγ δ Rαβγ δ
Now let us move on to actions containing modifications involving only Q = Rαβγ δ Rαβγ δ .
Our prototype example is f (Q) = −M 6/Q, with M another parameter with dimensions
of mass, and the analysis follows much the same as in the previous subsection, albeit with
different results.
Again, (16) demonstrates that there is a constant curvature vacuum solution to this action
given by
R(Q)
const = (24)1/3 M 2 .
(42)
What about other possible cosmological solutions? From (15), with the flat cosmological
ansatz, the analogue of the Friedmann equation becomes
M 6
24[ H 2 + 2H 2 H + 2H 4 ]3 h8H 8 + 36H 6 H + 54H 4 H 2
+ 20H 2 H 3 + 3 H 4 + 4H 5 H + 12H 3 H H + 6H H 2 H i = 0 .
3H 2 −
(43)
16
Employing our phase space technique once more, in the variables best-suited for analyzing
power-law behavior, the relevant equation is
x
dv
dx
1
2M 6 v [2v 2 − 6v + 3] (cid:2)−x6 (576v 6 − 1728v 5 + 2592v 4 − 2304v 3
= 2v +
+ 1296v 2 − 432v + 72) + M 6 (8v 6 − 36v 5 + 54v 4 − 20v 3 + 3v 2)(cid:3) .
It is clear from this that our action consisting of f (R, P , Q) = −M 6 /Q does not admit any
late-time power-law attractors.
(44)
2.5
2
1.5
v
1
0.5
-0.04
-0.03
-0.02
x
-0.01
0
0
FIG. 4: Phase portrait for the f (R, P , Q) = −M 6/Q modification in the coordinates (x, v), for
which an attractor at constant v = p corresponds to a power-law solution with a(t) ∝ tp .
This is consistent with a study of the late-time asymptotics of (44). As in our previous
analysis, late-time power-law solutions correspond to real solutions to the equation
0 + 62v 2
0 − 36v 3
8v 4
0 − 44v0 + 15 = 0 .
(45)
However, no such real solutions exist, confirming our phase-space analysis.
17
IV.
INCLUDING MATTER
We now show that the late-time behavior of our vacuum solutions remains unaltered upon
the inclusion of matter.
We begin by rewriting the equation of motion in a more convenient form. Let Σµν be the
tensor defined by the left hand side of (15). Then, the generalized Friedmann equation takes
the form
Σ00 = 8πGρ,
(46)
where ρ is the energy density of a perfect fluid with equation of state p = wρ. Now, since
x = −H , y = − x, and ρ ∝ a−3(1+w) , we have d ln(ρ/ρ0 )
dx
with (46) yields the equation of motion
= 3(1 + w) v
x . Combining this relation
x
dΣ00
dx
= 3(1 + w)vΣ00 .
(47)
Thus far, we have not imposed any dynamics. Specializing to a theory with f (R, P , Q) =
µ6/(aR2 + bP + cQ) gives
with
Σ00 =
1
24 x4 F (x, v , s),
F = 72 x6 + µ6 v 2 g (v ) − 2 x v h(v ) s
[d(v )]3
,
(48)
(49)
where s = dv
dx . The explicit forms of the functions h(v ), g (v ), and d(v ), defined in Appendix
A, are not needed in this section. Substituting (48) and (49) into (47) gives the equation
governing the dynamics of our theory in the presence of matter:
x
dF
dx
= γ (v ) F ,
where γ (v ) ≡ [3(1 + w) v + 4] . Using the chain rule this becomes
x(Fx + Fv s + Fs s′ ) = γ (v ) F ,
(50)
(51)
18
where
Fx =
Fv =
Fs =
∂F
∂x
∂F
∂v
∂F
∂s
= 432 x5 − 2µ6 v h s
d3
= µ6 v 2 gv + 2 v g − 2 x h s − 2 x v hv s
d3
2µ6 x v h
d3
= −
− 3 µ6 v 2g − 2xvhs
d4
dv
(52)
(53)
and a prime denotes differentiation with respect to x.
Seeking late-time power law behavior, we take the limits x → 0 and s → 0 in (51).
This yields the condition g (v0) = 0, which is solved by the same power law fixed points as
those obtained in vacuum. This is as one might expect, since matter redshifts away in the
asymptotic future. However, the above description proves useful for dealing with the issue of
stability. Since this is somewhat technical, we relegate the details to Appendix B and merely
assert here that these fixed points do remain stable in the presence of matter sources.
V. COMMENTS AND CONCLUSIONS
In the extreme low curvature regime, our only tests of general relativity are cosmological.
The discovery of new phenomena at these scales may point to new matter sources, but
alternatively may hint at hitherto undetected modifications of gravity.
The acceleration of the universe provides a particular challenge to modifications of gravity.
Unlike the known perturbative corrections to the Einstein-Hilbert action arising from string
theory, late time acceleration requires modifications that become important at extremely
low energies, so low that only today, at the largest scales in the universe, is the resulting
curvature low enough to lead to measurable deviations from general relativity.
These considerations led some of the current authors and others to consider new terms in
the action for gravity that consist of inverse powers of the Ricci scalar. It is easy to show that
such an approach introduces a de Sitter solution. However, this solution is unstable to a late-
time accelerating power law attractor. For appropriate choices of parameters, this theory
is a candidate to explain cosmic acceleration without the need for dark energy, although
the simplest such theories are in conflict with solar system tests. For Lagrangians that are
19
functions of only the Ricci scalar, there exists a map to an Einstein frame, in which the
new degrees of freedom are represented by a scalar field. As a result, such modified gravity
theories share many features in common with some dark energy models.
In this paper we have introduced a much more general class of modifications to the
Einstein-Hilbert action, becoming relevant at extremely low curvatures. Specifically, we
have considered inverse powers of arbitrary linear combinations of the curvature invariants
R2 , P ≡ Rµν Rµν and Q ≡ Rαβγ δ Rαβγ δ . Such modifications are not simply equivalent to
Einstein gravity plus scalar matter sources.
We have performed a general analysis of the late-time evolution of cosmological solutions
to these theories. Many of the theories exhibit late-time attractors of the form a(t) ∝ tp ,
with p some constant power. Indeed, there are often multiple such attractors. For a large
class of theories there exists at least one attractor satisfying p > 1, corresponding to cosmic
acceleration.
The detailed structure of cosmological solutions to these theories turns out to be quite
rich and varied, depending on the dimensionless parameters entering the particular linear
combination. Two distinct types of singularities may exist, as well as the late-time power
law attractors. We have identified all those theories for which the late time behavior is
consistent with the observed acceleration of the universe, providing a whole new class of
theories - generalized modified gravity theories - which are alternatives to dark energy.
The results we have found for the modification f = µ6/(a R2 + b P + c Q) may be gener-
alized to modifications
µ4n+2
(a R2 + b P + c Q)n ,
using exactly the same calculational techniques.
f =
(54)
In general there are power-law attractors with the following exponents, whenever they
8n2 + 10n + 2 − 3α ± √Γ
4(n + 1)
v gen
1,2 =
(55)
are real
,
where
Γ = 9n2α2 − (80n3 + 116n2 + 40n + 4) α
+ 64n4 + 160n3 + 132n2 + 40n + 4 .
(56)
20
Clearly, as n → ∞, the smaller attractor tends to 0, whereas the larger one increases linearly
as 4n.
Two special cases are
m4n+2
P n
M 4n+2
Qn
,
,
f = −
f = −
for which the power law attractors, v1,2 are
12n2 + 9n + 3 ± √144n2 + 120n3 − 15n2 − 30n − 3
2(3 + 3n)
4n2 + 2n + 1 ± √16n4 − 16n2 − 10n − 1
2n + 2
v (Q)
1,2 =
v (P )
1,2 =
.
(57)
(58)
(59)
For n = 1, all the above expressions agree with the values found earlier. It is interesting
to note that v (Q)
1,2 are imaginary for n = 1, but are real for all n > 1.
Of course, much remains to be done. We have not addressed solar system tests of these
theories since this is a complicated analysis that is beyond the scope of our current paper.
We have not focused on detailed comparisons between our models and the supernova data
and it is possible that there are specific signatures of this new physics in such data. For
example, the modified Friedmann equations arising from the theories presented here should
directly provide information about the jerk parameter. We intend to consider such effects,
not only in the present models, but in those proposed by us and other authors, in a separate
paper focused on the connections of these models with observations. In this context, as with
the case of the simple modifications introduced in [7], it may also be interesting to study
more complicated functions of the curvature invariants we have considered.
Acknowledgments
MT thanks Gia Dvali, Renata Kallosh, Burt Ovrut and Richard Woodard for helpful
discussions. The work of SMC is supported in part by the Department of Energy (DOE),
the NSF, and the Packard Foundation. VD is supported in part by the NSF and the DOE.
ADF, DAE and MT are supported in part by the NSF under grant PHY-0094122. DAE is
21
also supported in part by funds from Syracuse University and MT is supported in part by a
Cottrell Scholar Award from Research Corporation. MST is supported in part by the DOE
(at Chicago), the NASA (at Fermilab), and the NSF (at Chicago).
APPENDIX A: SOME DEFINITIONS
In Section III we defined
where
α =
4 (3a + b + c)
∆
,
∆ = 12a + 3b + 2c .
We should distinguish betweeen two cases:
1. ∆ = 0
In this case α diverges, but we may define
g (v ) = (3a + b + c) (282a + 69b + 44c) v + 15 (3a + b + c)2
h(v ) = 3 (3a + b + c)2
d(v ) = 3a + b + c .
(A1)
(A2)
(A3)
(A4)
(A5)
Note that both ∆, and 3a+b+c, cannnot vanish simultaneously, as this causes the Friedmann
equation to be singular for all values of x and v .
2. ∆ 6= 0
In this case we have
g (v ) = 2 ∆2 (v − s2 ) (v − s1) (v − p2 ) (v − p1)
α
4 (cid:17) (v − v2) (v − v1 )
h(v ) = ∆2 (cid:16)1 −
d(v ) = ∆ (v − p2) (v − p1) ,
22
(A6)
(A7)
(A8)
where either v1,2 or p1,2 are complex unless α = 1, in which case, v1,2 = p1,2 = s1 = 1/2 and
s2 = 3.75. Also s1,2 in general may be complex. As we saw in Section III, si , vi and pi are
all functions of α only.
APPENDIX B: STABILITY OF FIXED POINTS IN THE PRESENCE OF MAT-
TER
Let us rewrite (51) as a system of first-order ordinary differential equations
x′ = 1
γ (v ) F
x Fs −
v ′ = s .
s′ =
Fv
Fs
Fx
Fs
s −
(B1)
(B2)
(B3)
Treating x as a dependent variable makes the system autonomous, facilitating a phase-space
analysis.
To obtain the phase portrait of the system we define a vector field by ~W T ≡ (x′ , s′ , v ′ )
and plot this in the vicinity of the fixed point. In figure 5 we show a 2D slice of this phase
portrait by choosing a section at constant x. Note that at the fixed point, ~W (0) = (1, 0, 0).
To analyze the stability of the system, we first linearize the system about the fixed point
v = v0 , s = 0, and x = x0 ≪ 1 (all expressions evaluated at these values of v , s and x will
carry the superscript (0) )
We write the linearized system of equations as
Wi = Xα
the matrix M being defined by
α ) + W (0)
[M ](0)
iα (ηα − η (0)
i
,
and
,
∂ηα (cid:21)(0)
iα = (cid:20) ∂Wi
[M ](0)
x
23
~η ≡
.
s
v
(B4)
(B5)
(B6)
-0.1
-0.05
1E-6
5E-7
v
0
0
s
-5E-7
-1E-6
0.05
0.1
FIG. 5: Phase plot for the linearized equations close to the power law solution in the coordinates
(x, v), for which an attractor at constant v = p corresponds to a power-law solution with a(t) ∝ tp .
Now, introducing equilibrium coordinates
u1 = x − x0
u2 = s
u3 = v − v0 ,
we finally obtain the linearized equations
with σ0 = v0 gv,0
2 h0
.
u′
1 = 1
γ0 + σ0 − 1
x0
u′
2 =
u′
3 = u2 ,
u2 −
γ0 σ0
x2
0
u3
24
(B7)
(B8)
(B9)
(B10)
(B11)
(B12)
Notice that the first equation has decoupled from the other two. Hence, it suffices to
study the behavior of the subsystem (u2 , u3). We will use standard results from the theory
of dynamical systems to do so (see [38]).
In the vicinity of the fixed point, the behavior of the system can be classified by the
eigenvalues of the sub-matrix Mij , with i, j = 2, 3, with characteristic equation
2s λ − M (0)
λ2 − M (0)
2v = 0 .
(B13)
Since we are only interested in values of v and w (the equation of state parameter) which
satisfy v > 1/2 and −1 < w ≤ 1 (see [39, 40, 41] for arguments why this is a sensible
choice, and [42, 43] for how one may be tricked into inferring values outside this range
when considering gravitational theories other than General Relativity), we have that γ0 > 1.
Furthermore if we choose v0 = s2 (see Section III), the value of the bigger power law, typically
2s and M (0)
σ0 > 0. This implies that both M (0)
2v are negative, so that s2 is either a stable node
or a stable spiraling helix. The latter case is tantamount to stability.
We have a stable spiralizing helix if
2s )2 + 4 M (0)
(M (0)
2v < 0.
If not, we get a stable node. Relation (B14) leads to the condition
(γ0 + σ0 − 1)2 − 4 γ0 σ0 < 0 ,
(B14)
(B15)
which holds for all x0 ≪ 1.
Applying the last condition to the case α = 1 and w = 0 (dust), we have that (see
Appendix A)
and
σ0 =
4
3
s2 (s2 − 0.5) = 16.25
γ0 = 3 s2 + 4 = 15.25.
(B16)
(B17)
For these values of σ and γ , (B15) is easily satisfied, so that the power law solution s2 = 3.75
25
is a stable spiraling helix.
[1] A. G. Riess et al. [Supernova Search Team Collaboration], Astron. J. 116, 1009 (1998)
[arXiv:astro-ph/9805201 ].
[2] S. Perlmutter et al. [Supernova Cosmology Pro ject Collaboration], Astrophys. J. 517, 565
(1999) [arXiv:astro-ph/9812133 ].
[3] J. L. Tonry et al., arXiv:astro-ph/0305008 .
[4] C. L. Bennett et al., arXiv:astro-ph/0302207 .
[5] C. B. Netterfield et al.
[Boomerang Collaboration], Astrophys. J. 571, 604 (2002)
[arXiv:astro-ph/0104460 ].
[6] N. W. Halverson et al., Astrophys. J. 568, 38 (2002) [arXiv:astro-ph/0104489 ].
[7] S. M. Carroll, V. Duvvuri, M. Trodden and M. S. Turner, Phys. Rev. D 70, 043528 (2004)
[arXiv:astro-ph/0306438 ].
[8] S. Capozziello, S. Carloni and A. Troisi, arXiv:astro-ph/0303041 .
[9] C. Deffayet, G. R. Dvali and G. Gabadadze, Phys. Rev. D 65, 044023 (2002)
[arXiv:astro-ph/0105068 ].
[10] K. Freese and M. Lewis, Phys. Lett. B 540, 1 (2002) [arXiv:astro-ph/0201229 ].
[11] N. Arkani-Hamed, S. Dimopoulos, G. Dvali and G. Gabadadze, arXiv:hep-th/0209227 .
[12] G. Dvali and M. S. Turner, arXiv:astro-ph/0301510 .
[13] S. No jiri and S. D. Odintsov, Mod. Phys. Lett. A 19, 627 (2004) [arXiv:hep-th/0310045 ].
[14] N. Arkani-Hamed, H. C. Cheng, M. A. Luty and S. Mukohyama, arXiv:hep-th/0312099 .
[15] M. C. B. Abdalla, S. No jiri and S. D. Odintsov, arXiv:hep-th/0409177 .
[16] D. N. Vollick, Phys. Rev. D 68, 063510 (2003) [arXiv:astro-ph/0306630 ]
[17] R. Dick, Gen. Rel. Grav. 36, 217-224(2004) [arXiv:gr-qc/0307052 ]
[18] S. No jiri, S. D. Odintsov, Phys. Lett. B 576, 5-11(2003) [arXiv:hep-th/0307071 ]
[19] A. D. Dolgov, M. Kawasaki, Phys. Lett. B 573, 1-4(2003) [arXiv:astro-ph/0307285 ]
[20] S. No jiri, S. D. Odintsov, Phys. Rev. D 68, 123512(2003) [arXiv:hep-th/0307288 ]
26
[21] X. Meng, P. Wang, Class. Quant. Grav. 20, 4949-4962(2003) [arXiv:astro-ph/0307354 ]
[22] T. Chiba, arXiv:astro-ph/0307338 .
[23] X. Meng, P. Wang, arXiv:astro-ph/0308284
[24] E. E. Flanagan, Phys. Rev. Lett. 92, 071101(2004) [arXiv:astro-ph/0308111 ]
[25] M. E. Soussa, R. P. Woodard, Gen. Rel. Grav. 36, 855-862(2004) [arXiv: astro-ph/0308114 ]
[26] Y. Ezawa, H. Iwasaki, Y. Ohkuwa, N. Yamada, T. Yano, arXiv: gr-qc/0309010
[27] E. E. Flanagan, Class. Quant. Grav. 21, 417-426(2003) [arXiv:gr-qc/0309015 ]
[28] A. Ra jaraman, arXiv:astro-ph/0311160
[29] D. N. Vollick, Class. Quant. Grav. 21, 3813-3816 (2004) [arXiv:gr-qc/0312041 ]
[30] A. Nunez, S. Solganik, arXiv:hep-th/0403159
[31] G. Allemandi, A. Borowiec, M. Francaviglia, arXiv:hep-th/0407090
[32] A. Lue, R. Scoccimarro and G. Starkman, Phys. Rev. D 69,
044005 (2004)
[arXiv:astro-ph/0307034 ].
[33] A. Melchiorri, L. Mersini, C. J. Odman and M. Trodden, Phys. Rev. D 68, 043509 (2003)
[arXiv:astro-ph/0211522 ].
[34] D. N. Spergel et al., arXiv:astro-ph/0302209 .
[35] V. T. Gurovich and A. A. Starobinsky, Sov. Phys. JETP 50, 844 (1979) [Zh. Eksp. Teor. Fiz.
77, 1683 (1979)].
[36] A. A. Starobinsky, Phys. Lett. B 91, 99 (1980).
[37] We thank Gia Dvali, Renata Kallosh, Burt Ovrut and Richard Woodard for discussions on
this matter.
[38] S. H. Strogatz, “Nonlinear Dynamics and Chaos”, Addison Wesley, (1994).
[39] S. M. Carroll, M. Hoffman and M. Trodden, Phys. Rev. D 68, 023509 (2003)
[arXiv:astro-ph/0301273 ].
[40] J. M. Cline, S. y. Jeon and G. D. Moore, arXiv:hep-ph/0311312 .
[41] S. D. H. Hsu, A. Jenkins and M. B. Wise, arXiv:astro-ph/0406043 .
[42] S. M. Carroll, A. De Felice and M. Trodden, arXiv:astro-ph/0408081 .
[43] A. Lue and G. D. Starkman, arXiv:astro-ph/0408246 .
27
|
0811.0348 | 1 | 0811 | 2008-11-03T17:39:22 | Doppler maps and surface differential rotation of EI Eri from the MUSICOS 1998 observations | [
"astro-ph"
] | We present time-series Doppler images of the rapidly-rotating active binary star EI Eri from spectroscopic observations collected during the MUSICOS multi-site campaign in 1998, since the critical rotation period of 1.947 days makes it impossible to obtain time-resolved images from a single site. From the surface reconstructions a weak solar-type differential rotation, as well as a tiny poleward meridional flow are measured. | astro-ph | astro-ph |
Doppler maps and surface differential rotation of
EI Eri from the MUSICOS 1998 observations
Zs. Kovári∗, A. Washuettl†, B.H. Foing∗∗, K. Vida‡,∗, J. Bartus†, K. Oláh∗
and the MUSICOS 98 team§
∗Konkoly Observatory, Budapest, Hungary
†Astrophysikalisches Institut Potsdam, Germany
∗∗Research Support Division, ESA RSSD, ESTEC/SCI-SR, The Netherlands
‡Eötvös University, Department of Astronomy, Budapest, Hungary
§
Abstract. We present time-series Doppler images of the rapidly-rotating active binary star EI Eri
from spectroscopic observations collected during the MUSICOS multi-site campaign in 1998, since
the critical rotation period of 1.947 days makes it impossible to obtain time-resolved images from a
single site. From the surface reconstructions a weak solar-type differential rotation, as well as a tiny
poleward meridional flow are measured.
Keywords: Stars: activity, Stars: late-type, Stars: imaging, Starspots, Stars: individual: EI Eri
PACS: 97.10.Jb, 97.10.Qh ,97.20.Jg
INTRODUCTION
Due to the fast disperse of high resolution spectroscopic observational facilities, surface
differential rotation (hereafter DR) measurements has become achievable for stars of
different types, giving a useful contribution for developing solar and stellar dynamo
theory. Indirect imaging techniques, such as Doppler imaging can help in extending
this input knowledge by tracing the positions of different spots/magnetic features on
consecutive surface maps. The method of cross-correlating consecutive Doppler images
to follow the time evolution of the spots and to derive surface DR was demonstrated for
the first time for AB Dor [1], and it has been used extensively thenceforth. Unfortunately,
rapid spot changes can fade out the correlation pattern of the DR. On the other hand,
averaging many cross-correlation maps can overwhelm the masking effect of random-
like short-term spot changes [2].
In this paper we present time-series Doppler images for the rapidly rotating active
binary star EI Eridani (HD 26337). From the consecutive surface maps we measure
surface DR. Since the rotation period of the star is nearly 2 days, data with sufficient
phase coverage for a Doppler image can only be collected over about 15 days from a
single observing site. Thus, the reconstructed surface map is necessarily time-averaged,
which makes DR measurements unfeasible. To avoid this problem, in our study we use
a set of high resolution spectra collected in the fifth MUSICOS (MUlti-SIte COntinuous
Spectroscopy) campaign in 1998, which allows to derive time-resolved surface maps.
FIGURE 1. Observing sites involved in the MUSICOS 98 campaign
OBSERVATIONS
The fifth MUSICOS campaign took place from November 22 to December 11, 1998.
It involved eight northern and southern sites and ten telescopes of 2m class, mostly
equipped with cross-dispersed echelle spectrographs. Fig. 1 shows the distribution of
the participating observatories. From the total of 122 high-resolution spectra obtained
in the campaign for EI Eridani, 81 spectra were suitable for the purposes of Doppler
imaging. Spectroscopic observations are supported by simultaneous photometry carried
out with the Wolfgang-Amadeus twin Automatic Photoelectric Telescope (APT) at
Fairborn Observatory, Arizona [3]. Phases of all line profiles are computed using the
ephemeris HJD = 2 448 054.7130 + 1.9472287 × E [4].
TIME-SERIES DOPPLER IMAGING
From the chronologically listed 81 spectra obtained during 20 days, we form 52 subsets
with 30-30 spectra. Each subset is formed from the preceeding one by dropping the
spectrum from the beginning of the list and adding the subsequent one to the end. From
these overlapping subsets we compute 52 surface maps using our Doppler imaging code
TempMap [5]. As astrophysical input we use the data listed in Table 1 taken from [6].
The resulted Doppler maps confirm the existence of a stable polar spot changing slightly
in size and shape, while low latitute spots are found to be short lived. Independent
inversions for the CaI-6439 and for the FeI-6411 mapping lines are in good agreement,
as seen in the corresponding example image pair in Fig. 2.
DIFFERENTIAL ROTATION AND MERIDIONAL FLOW
For measuring surface DR we apply the technique called ACCORD (Average Cross-
CORrelation of time-series Doppler images) introduced for the time-series Doppler
maps of LQ Hya in [2], and applied more recently to s Gem [7] and to z And [8]. The 52
time-series Doppler images are coupled in order to obtain 22 pairs of non-overlapping,
TABLE 1. Stellar parameters of EI Eri
Parameter
Value
Pphot = Prot
T0,phot =T0,rot
g
K1
e (eccentricity)
Tphot
Tmax
vsin i
Inclination i
logg
Microturbulence x
Macroturbulence z r = z t
log[Ca] abundance
log[Fe] abundance
1.9472324 d
2448054.7109d
21.64 km/s
26.83 km/s
0 (adopted)
5500 K
6000 K
51.0 km/s
56.0◦
3.5
2.0 km/s
4.0 km/s
−6.3 (0.6 dex below solar)
−5.5 (1.1 dex below solar)
FIGURE 2. Simultaneous example image reconstructions from the time-series for the CaI-6439 (top)
and for the FeI-6411 (bottom) mapping lines
consecutive Doppler maps. These pairs are cross-correlated and the resulting correlation
function (ccf) maps are averaged. On the resulting average ccf map, we fit the correlation
peak for each latitudinal strip with a Gaussian profile. The Gaussian peaks per latitude
strips then represent the DR pattern and can thus be fitted with a standard solar-like
quadratic DR law. The best fit DR function for the combined Ca+Fe ccf map in the left
panel of Fig. 3 indicates a DR parameter a = D
of 0.037, about one-sixth of the
value measured on the Sun.
/W
An average meridional flow can be quantified by using ACCORD, since latitudinal
motion of spots can be analysed by cross-correlating the corresponding longitude strips
along the meridian circles. For this we use only the hemisphere of the visible pole where
Doppler imaging is more reliable. The correlation maxima for each longitudinal strip are
fitted with a Gaussian and the resulting peaks represent the best correlating latitudinal
shifts. For a detailed description of the method see [7]. The latitudinal correlation pattern
plotted in the right panel of Fig. 3 can be converted into an average poleward velocity
field of less than 100 m/s. For further details see our forthcoming paper [9].
W
t
f
i
h
s
e
d
u
t
i
t
a
L
20
10
0
-10
-20
0.900
0.800
0.700
0.600
0.500
0.400
0.300
0.200
0.100
FIGURE 3. Resulting average ccf map with the best fit DR law (left) and the latitudinal ccf map (right)
0
50
100
150
200
Longitude
250
300
350
ACKNOWLEDGMENTS
ZsK is a grantee of the Bolyai János Scholarship of the Hungarian Academy of Sciences.
ZsK, AW, KV and KO are supported by the Hungarian Science Research Program
(OTKA) grants T-048961 and K 68626.
REFERENCES
1. J.-F. Donati, and A. Collier Cameron, MNRAS 291, 1 (1997).
2. Z. Kovári, K. G. Strassmeier, T. Granzer, M. Weber, K. Oláh, and J. B. Rice, A&A 417, 1047 (2004).
3. K. G. Strassmeier, L. J. Boyd, D. H. Epand, and T. Granzer, PASP 109, 697 (1997).
4. A. Washuettl, K. G. Strassmeier, B. Foing, and MUSICOS98 Team, "MUSICOS Observations of the
Chromospherically Active Binary Star EI Eridani," in 12th Cambridge Workshop on Cool Stars, Stellar
Systems, and the Sun, eds. A. Brown, G.M. Harper, and T.R. Ayres, (University of Colorado), 2003, p.
1008.
5. J. B. Rice, W. H. Wehlau, and V. L. Khokhlova, A&A 208, 179 (1989).
6. A. Washuettl, K. G. Strassmeier, and M. Weber, Astron. Nachr. (2008), submitted.
7. Z. Kovári, J. Bartus, K. G. Strassmeier, K. Vida, M. Švanda, and K. Oláh, A&A 474, 165 (2007).
8. Z. Kovári, J. Bartus, K. G. Strassmeier, K. Oláh, M. Weber, J. B. Rice, and A. Washuettl, A&A 463,
1071 (2007).
9. A. Washuettl, Z. Kovári, B. H. Foing, K. Oláh, M. Weber, D. García-Alvarez, O. J., Y. Unruh, and
MUSICOS 98 Team, A&A (2008), submitted.
|
astro-ph/0608443 | 1 | 0608 | 2006-08-21T19:46:08 | Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data | [
"astro-ph"
] | We propose a useful method for mapping large-scale velocity fields in the solar photosphere. It is based on the local correlation tracking algorithm when tracing supergranules in full-disc dopplergrams. The method was developed using synthetic data. The data processing the data are transformed during the data processing into a suitable coordinate system, the noise is removed, and finally the velocity field is calculated. Resulting velocities are compared with the model velocities and the calibration is done. From our results it becomes clear that this method could be applied to full-disc dopplergrams acquired by the Michelson Doppler Imager (MDI) onboard the Solar and Heliospheric Observatory (SoHO). | astro-ph | astro-ph |
Astronomy&Astrophysicsmanuscript no. aa4124-05
(DOI: will be inserted by hand later)
November 1, 2018
Large-scale horizontal flows in the solar photosphere I
Method and tests on synthetic data
M. Svanda1,2, M. Klvana1, and M. Sobotka1
1 Astronomical Institute, Academy of Sciences of the Czech Republic, Fricova 298, CZ-251 65, Ondrejov, Czech Republic
e-mail: [email protected], [email protected], [email protected]
2 Astronomical Institute, Charles University in Prague, V Holesovick´ach 2, CZ-180 00, Prague 8, Czech Republic
Received ; accepted
Abstract. We propose a useful method for mapping large-scale velocity fields in the solar photosphere. It is based on the
local correlation tracking algorithm when tracing supergranules in full-disc dopplergrams. The method was developed using
synthetic data. The data processing the data are transformed during the data processing into a suitable coordinate system, the
noise is removed, and finally the velocity field is calculated. Resulting velocities are compared with the model velocities and
the calibration is done. From our results it becomes clear that this method could be applied to full-disc dopplergrams acquired
by the Michelson Doppler Imager (MDI) onboard the Solar and Heliospheric Observatory (SoHO).
Key words. Sun: photosphere -- Methods: data analysis
1. Introduction
The solar photosphere is a very dynamic layer of the solar
atmosphere. It is strongly influenced by the underlying con-
vection zone. Despite years of intensive studies, the velocity
fields in the solar photosphere remain not very well known.
The evidence of the vigorous and sometimes chaotic charac-
ter of the motions of observed structures in the photosphere
(sunspots, granules, and other features) came already from the
first systematic studies made in 19th century, of which let us
at least mention the discovery of the solar differential rotation
(Carrington 1859). Motions in the photosphere are strongly
coupled with magnetic fields. The large-scale velocity fields
are very important for studies of the global solar dynamo.
An attempt
to describe the differential
rotation by
results.
a parabolic dependence did not make for clear
Coefficients of the parabola differed according to traced objects
and also changed with time, when tracing one type of object
(reviewed e. g. by Schroter 1985). The character of the differ-
ential rotation has never been in doubt.
It follows from these arguments that a temporally vari-
able streaming of the plasma exists on the surface of the Sun,
which can be roughly described by the differential rotation.
This streaming has a large-scale character, large-scale plasma
motions were studied for example on the basis of tracking the
magnetic structures (e. g. Ambroz 2001a, 2001b). The long-
term Doppler measurements done by the MDI onboard SoHO
Send offprint requests to: M. Svanda
make it possible to extend the studies of large-scale velocities
in the solar photosphere. The knowledge of the behaviour of
velocities in various periods of the solar activity cycle could
contribute to understanding the coupling between the velocity
and magnetic fields and of the solar dynamo function.
There are at least three methods calculating photospheric
velocities:
1. Direct Doppler measurement -- provides only one compo-
nent (line-of-sight) of the velocity vector. These velocities
are generated by local photospheric structures, amplitudes
of which are significantly greater than amplitudes of the
large-scale velocities. The complex topology of such struc-
tures complicates an utilisation for our purpose. Analysing
this component in different parts of the solar disc led to
very important discoveries (e. g. supergranulation -- Hart
1956 and Leighton et al. 1962).
2. Tracer-type measurement -- provides two components of
the velocity vector. When tracing some photospheric trac-
ers, we can compute the local horizontal velocity vectors in
the solar photosphere. Tracking motions of sunspots across
the solar disc led to the discovery of the differential rotation
(Carrington 1859).
3. Local helioseismology -- provides a full velocity vector.
Although local helioseismology (e. g. Zhao 2004) is a very
promising method, it is still in progress and until now does
not provide enough reliable results.
2
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
Fig. 1. Comparison of the real dopplergram (left, observed by MDI/SoHO) and the simulated one (right). Both images are visually
very similar. The black colour means line-of-sight velocity −700 m s−1 (towards an observer) while white represents +700 m s−1.
Since the photosphere is a very thin layer (0.04 % of the
solar radius), the large-scale photospheric velocity fields have
to be almost horizontal. Then, the tracer-type measurement
should be sufficient for mapping the behaviour of such veloci-
ties. In this field the local correlation tracking (LCT) method is
very useful.
This method was originally designed for the removal of
the seeing-induced distortions in image sequences (November
1986) and later used for mapping the motions of granules in the
series of white-light images (November & Simon 1988). The
method works on the principle of the best match of two frames
that record the tracked structures at two different instants. For
each pixel in the first frame, a small correlation window is cho-
sen and is compared with a somewhat displaced window of the
same size in the second frame. A vector of displacement is then
defined as a difference in the coordinates of the centres of both
windows when the best match is found. The velocity vector is
calculated from this displacement and the time lag between two
frames.
LCT was recently used for tracking many features in vari-
ous types of observations, especially for tracking the granules
in high-resolution white-light images (e. g. Sobotka et al. 1999,
2000).
The method needs a tracer -- a significant structure recorded
in both frames, the lifetime of which is much longer than the
time lag between the correlated frames. We decided to use
the supergranulation pattern in the full-disc dopplergrams, ac-
quired by the MDI onboard SoHO. We assume that supergran-
ules are carried as objects by the large-scale velocity field. This
velocity field is probably located beneath the photosphere, so
that the resulting velocities will describe the dynamics in both
the photospheric and subphotospheric layer. The existence of
the supergranulation on almost the whole solar disc (in contrast
to magnetic structures) and its large temporal stability make the
supergranulation an excellent tracer.
The resulting velocities cover the whole solar disc. Hence
the velocity field also describes a motion of the supergranu-
lation in the areas occupied by active regions or by magnetic
field concentrations. This fact allows the results to be used to
study the mutual motions of substructures like sunspots, mag-
netic field in active regions, background fields, and the quiet
photosphere.
Supergranules are structures with a strong convection cou-
pling. The mean size of the supergranular cell is approx. 30 Mm
(e. g. Wang & Zirin 1989), with the size and shape depen-
dent on the phase of the solar cycle (e. g. S´ykora 1970).
Supergranules are quite stable with a mean lifetime of approx.
20 hours (e. g. Leighton 1964). Information about the distri-
bution function of the lifetime was published by DeRosa et
al. (2000). The internal velocity field in the supergranular cell
is predominantly horizontal (e. g. Hathaway et al. 2002) with
the amplitude approx. 300 m s−1. Due to the horizontality of
the internal velocity field, the supergranules can be observed in
dopplergrams on the whole solar disc except for its centre.
We do not propose that this method capable of measuring
velocities of order 1 m s−1, but we do expect that the large-
scale velocities will have magnitudes at least one order greater.
We also have to take the largest-scale velocities into account
like the differential rotation or meridional circulation, which
have velocities of at least 10 m s−1. If for example we take
differential rotation described by the formula ω[deg/day] =
13.064 − 1.69 sin2 b − 2.35 sin4 b (Snodgrass 1984), we have
to expect a velocity approx. 190 m s−1 in b = 60 ◦ in the
Carrington's coordinate system. The main goal of our study
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
3
(and the proposed method of mapping the velocities should be
proxy for this purpose) is to separate the superposed velocity
field into the components and to investigate their physics.
This paper is the first one of a series about the large-scale
photospheric velocity fields. In the following papers we shall
apply the method here suggested to observed data and describe
the properties of the large-scale velocity fields in different pe-
riods of the solar activity cycle.
2. Synthetic data
A recent experience with applying this method to observed
data (e. g. Svanda et al. 2005) has shown that for the proper
setting of the parameters and for the tuning of the method,
synthetic (model) data with known properties are needed. The
synthetic data for the analysis come from a simple simulation
(SISOID code = SImulated Supergranulation as Observed In
Dopplergrams) with the help of which we can reproduce the
supergranulation pattern in full-disc dopplergrams.
The SISOID code is not based on physical principles taking
place in the origin and evolution of supergranulation, but in-
stead on a reproduction of known parameters that describe the
supergranulation. Individual synthetic supergranules are char-
acterised as centrally symmetric features described by their po-
sition, lifetime (randomly selected according to the measured
distribution function of the supergranular lifetime -- DeRosa et
al. 2000), maximal diameter (randomly according to its distri-
bution function -- Wang & Zirin 1989) and characteristic values
of their internal horizontal and vertical velocity components
(randomly according to their distribution function -- Hathaway
et al. 2002).
The most important simplification in the SISOID code is
that individual supergranules do not influence each other, but
simply overlap. The final line-of-sight velocity at a certain
point is given by the sum of line-of-sight velocities of indi-
vidual synthetic supergranules at the same position.
New supergranules can arise inside the triangle of neigh-
bouring supergranules (identification of such triangles is done
by the Delaunay triangulation algorithmed by Barry 1991) only
when the triangle is not fully covered by other supergranules
and when anyone of the supergranules located at the vertices of
the triangle is not too young, so that in the future it could fully
cover the triangle. The position of the origin of the new super-
granule is the centroid of the triangle; each vertex is weighted
by the size of its supergranule.
The SISOID simulation is done in the pseudocylindri-
cal Sanson-Flamsteed coordinate grid (Calabretta & Greisen
2002); the transformation from the heliographic coordinates is
given by the formula:
x = ϑ cos ϕ, y = ϕ,
(1)
where x and y are coordinates in the Sanson-Flamsteed coor-
dinate system, and ϑ and ϕ are heliographic coordinates origi-
nating at the centre of the disc. At each step an appropriate part
of the simulated supergranular field is transformed into helio-
graphic coordinates. The output of the program is a synthetic
dopplergram of the solar hemisphere in the orthographical pro-
jection to the disc. We assume in our simulation that the Sun
lies in an infinite distance from the observer and that the P (po-
sition angle of solar rotation axis) and b0 (heliograhic latitude
of centre of solar disc) angles are known.
Each step in calculation includes the evaluation of the pa-
rameters of individual supergranules, then small old supergran-
ules under the threshold (2 Mm in size) are removed from the
simulation and all the triangles are checked, whether a new su-
pergranule can arise inside them. This step in the SISOID code
corresponds to 5 minutes in real solar time. The computation is
always started from the regular grid. Properties of "supergran-
ules" are chosen randomly according to their real distribution
functions. The first 1000 steps are "dummy", i. e., no vector
velocity field is included and no synthetic dopplergram is cal-
culated. This starting interval is taken for the stabilisation of the
supergranular pattern. In the next steps, the model vector veloc-
ity field has already been introduced. This field influences only
the positions of individual cells. The dopplergram is calculated
every third step. For one day in real solar time, 96 dopplergrams
are calculated.
The model velocity field with Carrington rotation added is
applied according to the assumption of the velocity analysis
that the supergranules are carried by a velocity field on a larger
scale. In the simulation only, the position of individual super-
granules is influenced, and no other phenomena are taken into
account. These synthetic dopplergrams are visually similar to
the real observed dopplergrams (see Fig. 1).
3. Method of data processing
The MDI onboard SoHO acquired the full-disc dopplergrams at
a high cadence in certain periods of its operation -- one obser-
vation per minute. These campaigns were originally designed
for studying the high-frequency oscillations. The primary data
contain lots of disturbing effects that have to be removed be-
fore ongoing processing: the rotation line-of-sight profile, p-
modes of solar oscillations. We detected some instrumental ef-
fects connected to the data-tranfer errors. It is also known that
the calibration of the MDI dopplergrams is not optimal (e. g.
Hathaway et al. 2002) and has to be corrected to avoid system-
atic errors. While examining long-term series of MDI dopp-
lergrams, we have met systematic errors connected to the re-
tuning of the interferometer. We should also take those geo-
metrical effects into account (finite observing distance of the
Sun, etc.) causing bias in velocity determination. According
to Strous (2000) for example, the bias coming out of a per-
spective is about 2 m s−1, and it depends on the position on
the disc. It has been proven (e. g. Liu & Norton 2001) that
MDI provides reliable velocity measurements when the mag-
netic field is lower than 2000 Gauss. The velocity observation
by MDI will induce up to 100 % error if the magnetic field is
higher than 3000 Gauss due to the magnetic sensitivity of the
used Ni I line and the limitations of computational algorithm,
which cause crosstalks between measured MDI dopplergrams
and magnetograms. The removal of these effects will be de-
scribed in detail in the next paper, while the synthetic data used
in this study do not suffer from these phenomena.
As input to the data processing we take a one-day observa-
tion that contains 96 full-disc dopplergrams in 15-minute sam-
4
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
Fig. 2. Left -- model vector velocity field. Right -- a velocity field that was computed by applying the LCT method to the synthetic
supergranulation pattern with the imposed model field. The arrow lengths, representing the velocity magnitudes, have the same
scale. The images are visually very similar, however the magnitudes of calculated velocities are underestimated.
pling. Structures in these dopplergrams are shifted with respect
to each other by the rotation of the Sun and by the velocity field
under study.
First, the shift caused by the rotation has to be removed. For
this reason, the whole data series (96 frames) is "derotated"
using Carrington's rotation rate, so that the heliographic lon-
gitude of the central meridian is equal in all frames and also
equals the heliographic longitude of the central meridian of the
central frame of the series. This data-processing step causes the
central disc area ("blind spot" caused by prevailing horizontal
velocity component in supergranules) in the derotated series to
move with the Carrington's rate. During the "derotation" the
seasonal tilt of the rotation axis towards the observer (given
by b0 -- heliographic latitude of the centre of the disc) is also
removed, so that b0 = 0 in all frames.
Then the data series is transformed to the Sanson-
Flamsteed coordinate system to remove the geometrical dis-
tortions caused by the projection of the sphere to the disc.
Parallels in the Sanson-Flamsteed pseudocylindrical coordinate
system are equispaced and projected at their true length, which
makes it an equal area projection. Formulae of the transforma-
tion from heliographic coordinates are given by Eq. (1).
The noise coming from the evolutionary changes in the
shape of individual supergranules and the motion of the "blind
spot" in the data series with the Carrington's rotational rate
are suppressed by the k-ω filter in the Fourier domain (Title
et al. 1989, Hirzberger et al. 1997). The cut-off velocity is set
to 1 500 m s−1 and has been chosen on the basis of empirical
experience.
The existence of the differential rotation complicates the
tracking of the large-scale velocity field, because the ampli-
tudes and directions of velocities of the processed velocity field
have a significant dispersion. We have found that, when the
scatter of magnitudes is too large, velocities of several hun-
dred m s−1 cannot be measured precisely by the LCT algorithm
where the displacement limit for correlation was set to detect
velocities of several tens of m s−1. Therefore the final velocities
are computed in two steps. The first step provides a rough in-
formation about the average zonal flows using the differential
rotation curve
ω = A + B sin2 b + C sin4 b ,
and calculating its coefficients.
(2)
In the second step this average zonal flow is removed from
the data series, so that during the "derotation" of the whole
series the differential rotation inferred in the first step and ex-
pressed by (2) is used instead of the Carrington rotation. The
scatter of the magnitudes of the motions of supergranules in the
data transformed this way is much smaller, and a more sensitive
and precise tracking procedure can be used.
The LCT method is used in both steps. In the first step, the
checked range of velocity magnitudes is set to 200 m s−1, but
the accuracy of the calculated velocities is roughly 40 m s−1.
In the second step the range is only 100 m s−1 with much bet-
ter accuracy. The lag between correlated frames equals in both
cases 16 frame intervals (i. e. 4 hours in solar time), and the
correlation window with FWHM 30 pixels equals 60′′ on the
solar disc in the linear scale. In one observational day, 80 pairs
of velocity maps are calculated and averaged.
For
the
calculation we use
adapted program
flowmaker.pro originally written in IDL by Molowny-
Horas & Yi (1994). The algorithm has a limitation in the range
the
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
5
of displacements that are checked for each pixel. The quality of
correspondence (in our case the sum of absolute differences of
both correlation windows) is computed in nine discrete points,
then the biquadratic surface is fitted through these nine points,
and an extremum position (Darvann 1991) is calculated. The
final displacement vector is equivalent to the position of the
extremum.
4. Results
In our tests we have used lots of variations of simple axisym-
metric model flows (with a wide range of values of parame-
ters describing the differential rotation and meridional circula-
tion) with good success in reproducing the models. When com-
paring the resulting vectors of motions with the model ones,
we found a systematic offset in the zonal component equal to
voffset, zonal = −15 m s−1. This constant offset appeared in all the
tested model velocity fields and comes from the numerical er-
rors during the "derotation" of the whole time series. For the
final testing, we used one of the velocity fields obtained in our
previous work ( Svanda et al. 2005). This field approximates the
velocity distribution that we may expect to observe on the Sun.
The model flows have structures with a typical size of 60′′,
since they were obtained with the correlation window of this
size.
The calculated velocities (with voffset, zonal = −15 m s−1
corrected) were compared with the model velocities (Fig. 2).
Already from the visual impression it becomes clear that
most of vectors are reproduced very well in the direction,
but the magnitudes of the vectors are not reproduced so well.
Moreover, it seems that the magnitudes of vectors are underes-
timated. This observation is confirmed when plotting the mag-
nitudes of the model vectors versus the magnitudes of the cal-
culated vectors (Fig. 3). The scatter plot contains more than
1 million points, and most of the points concentrate along a
strong linear dependence, which is clearly visible. This depen-
dence can be fitted by a straight line that can be used to derive
the calibration curve of the magnitude of calculated velocity
vectors. The calibration curve is given by the formula
vcor = 1.13 vcalc,
(3)
where vcalc is the magnitude of velocities coming from the LCT,
and vcor the corrected magnitude. The directions of the vectors
before and after the correction are the same. The uncertainty of
the fit can be described by 1-σ-error 15 m s−1 for the velocity
magnitudes under 100 m s−1 and 25 m s−1 for velocity mag-
nitudes greater than 100 m s−1. The uncertainties of approx.
15 m s−1 have their main origin in the evolution of supergran-
ules. We studied the dependence of the error of velocity deter-
mination on the time lag used when no model velocity field was
introduced. We found that this dependence is slowly increasing
with the time lag (Fig. 4) due to the evolution of individual
supergranules. Evolution of supergranules is only one part of
story, but it gives the lower limit of accuracy that can be ob-
tained by this method. We also ran a test of the LCT sensitivity
on the evolution of supergranules when a known underlying
velocity field is introduced and came to similar results.
Fig. 3. Scatter plot for inference of the calibration curve.
Magnitudes of calculated velocities are slightly underestimated
by LCT, but the linear behaviour is clearly visible. A line rep-
resenting the 1:1 ratio is displayed. The calibration affects only
the magnitudes of the flows, while the directions do not need
any correction.
Fig. 4. Dependence of the 1-σ-error of the calculated velocity
on the time lag when no model velocity field was introduced
and only the evolution of supergranules have been taken into
account. The dashed line represents lag 16 (4 hours) that is usu-
ally used in our method.
We tested the sensitivity of the method to the choice of val-
ues of FWHM of the correlation window and of the lag be-
tween correlated frames in the LCT method. We found that our
method is practically insensitive to the choice of the time lag
between correlated frames when the lag is in the interval of 10 --
24 (2.5 -- 6 hours). The larger the lag we choose, the lower the
velocities we are able to detect. On the other hand, we have to
take into account that a larger lag between correlated frames
causes more noise in calculated results coming from evolution-
ary changes of supergranules and probably also from evolution-
ary changes in the velocity field under study. According to our
tests, for a time lag greater than 30 (7.5 hours), the numerical
6
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
Fig. 5. Influence of the calculated velocity field on the choice of FWHM of the correlation window. Left -- 120′′, right -- 200′′.
The model velocity field is the same as in Fig. 2.
noise raises very fast. The lag 16 (4 hours) seems to be a good
tradeoff between sensitivity and noise.
The choice of different FWHMs of the correlation window
changes the spatial resolution according to FWHM. The gen-
eral character of the vector field is preserved within the limits of
resolution (cf. Fig. 5). Larger FWHM causes a smoothing of re-
sults and an underestimation of vector magnitudes (cf. Fig. 6).
We found the used parameters (FWHM 60′′, lag 4 hours) to
be the best compromise, however these values can be changed
during the work on real data.
5. Conclusions
We have developed a method for mapping the velocity fields
in the solar photosphere based on the local correlation track-
ing (LCT) algorithm. The method consists in two main steps.
In the first main step the mean zonal velocities are calculated
and, on the basis of expansion to the Fay's formula (2), the
differential rotation is removed. In the second main step, the
LCT algorithm with an enhanced sensitivity is applied. Finally,
the differential rotation (obtained in the first step) is added to
the vector velocity field obtained in the second main step. Both
main steps can be divided into a few substeps, which are mostly
common.
1. "Derotation" of the data series using the Carrington's rota-
tion rate in the first step and using the calculated differential
rotation in the second step.
2. Transformation of
the data series into the Sanson-
Flamsteed coordinate system to remove the geometrical
distortion caused by the projection to the disc.
3. The k-ω filtering (with the cut-off velocity 1 500 m s−1)
for suppression of the noise coming from the evolutionary
changes of supergranules, of the numerical noise, and for
the partial removal of the "blind spot" (an effect at the cen-
tre of the disc).
4. Application of the LCT: the lag between correlated frames
is 4 hours, the correlation window with FWHM 60′′, the
measure of correlation is the sum of absolute differences
and the nine-point method for calculation of the subpixel
value of displacement is used. The calculated velocity field
is averaged over the period of one day.
Fig. 6. Histograms of velocity magnitudes for various FWHM
of the LCT algorithm.
Finally, the magnitude of the calculated vector field is corrected
using formula (3) obtained from the tests on the synthetic data.
M. Svanda et al.: Large-scale horizontal flows in the solar photosphere I: Method and tests on synthetic data
7
According to our test, the method provides a very reliable
tool for mapping the velocity fields with the spatial resolution
of 60′′ (43,5 Mm) on the solar disc and with the accuracy of
15 m s−1 for velocity magnitudes under 100 m s−1 and 25 m s−1
for velocity magnitudes greater than 100 m s−1. The method is
ready to use on the real full-disc dopplergrams acquired by the
MDI onboard SoHO. Our method shows that the usability of
the LCT method is wider than originally expected.
Acknowledgements. The authors of this paper were supported by
the Czech Science Foundation under grants 205/03/H144 (M. S.)
and 205/04/2129 (M. K.), by the Grant Agency of the Academy of
Sciences of the Czech Republic under grant IAA 3003404 (M. S.)
and by ESA-PECS under grant No. 8030 (M. S.). The Astronomical
Institute is working on the Research project AV0Z10030501 of the
Academy of Sciences of the Czech Republic. The MDI data were
kindly provided by the SoHO/MDI consortium. SoHO is the project
of international cooperation between ESA and NASA.
References
Ambroz, P. 2001a, Sol. Phys., 198, 253
Ambroz, P. 2001b, Sol. Phys., 199, 251
Barry J. 1991, Advances in Engineering Software, 13, 325
Carrington, R. C. 1859, MNRAS, 19, 81
Calabretta, M. R., & Greissen, E. W. 2002, A&A, 395, 1077
Darvann, T. A. 1991, Ph.D. Thesis, University of Oslo
DeRosa, M. L., Lisle, J. P., & Toomre, J. 2000, SPD Meeting at Lake
Tahoe, article 1.06
Hart, A. B. 1956, MNRAS, 116, 38
Hathaway, D. H., Beck, J. G., Han, S., & Raymond, J. 2002,
Sol. Phys., 205, 25
Hirzberger, J., V´azquez, M., Bonet, J. A., Hanslmeier, A., & Sobotka,
M. 1997, ApJ, 480, 406
Leighton, R. B., Noyes, R. W., & Simon, G. W. 1962, ApJ, 135, 474
Leighton, R. B. 1964, ApJ, 140, 1547
Liu, Y., & Norton, A. A. 2001, SOI Technical Note 01-144, HEPL,
Stanford University, 35 pages
Molowny-Horas, R., & Yi, Z. 1994, Internal Report No. 31, Institute
of Theoretical Astrophysics, University of Oslo
November, L. J. 1986, Appl. Opt., 25, 392
November, L. J., & Simon, G. W. 1988, ApJ, 333, 427
Schroter, E. H. 1985, Sol. Phys., 100, 141
Snodgrass, H. M. 1984, Sol. Phys., 94, 13
Sobotka, M., V´azquez, M., Bonet, J. A., Hanslmeier, A., & Hirzberger,
J. 1999, ApJ, 511, 436
Sobotka, M., V´azquez, M., Cuberes, M. S., Bonet, J. A., &
Hanslmeier, A. 2000, ApJ, 544, 1155
Strous, L. H. 2000, Sol. Phys., 195, 219
Svanda, M., Klvana M., & Sobotka, M. 2005, Hvar Obs. Bull., 29, 39
S´ykora, J. 1970, Sol. Phys., 13, 292
Title, A. M., Tarbell, T. D., Topka, K. P., Ferguson, S. H., Shine, R.
A., et al. 1989, ApJ, 336, 475
Wang, H., & Zirin, H. 1989, Sol. Phys., 120, 1
Zhao, J. 2004, Ph.D. Thesis, Stanford University
|
astro-ph/0205300 | 1 | 0205 | 2002-05-17T15:27:36 | Rotation and activity in the solar-metallicity open cluster NGC2516 | [
"astro-ph"
] | We report new measures of radial velocities and rotation rates (v sin i) for 51 F and early-G stars in the open cluster NGC2516, and combine these with previously published data. From high signal-to-noise spectra of two stars, we show that NGC2516 has a relative iron abundance with respect to the Pleiades of delta([Fe/H])= +0.04 +/- 0.07 at the canonical reddening of E(B - V) = 0.12, in contrast to previous photometric studies that placed the cluster 0.2 to 0.4 dex below solar. We construct a color-magnitude diagram based on radial velocity members, and explore the sensitivity of photometric determinations of the metallicity and distance to assumed values of the reddening. For a metal abundance near solar, the Hipparcos distance to NGC2516 is probably underestimated. Finally, we show that the distribution of rotation rates and X-ray emission does not differ greatly from that of the Pleiades, when allowance is made for the somewhat older age of NGC2516. | astro-ph | astro-ph |
Rotation and activity in the solar-metallicity open cluster
NGC 25161,2
Donald M. Terndrup3 and Marc Pinsonneault
Department of Astronomy, Ohio State University, 140 West 18th Avenue, Columbus, OH
43210
[email protected], [email protected]
Robin D. Jeffries and Alison Ford4
Department of Physics, Keele University, Keele, Staffordshire ST5 5BG, United Kingdom
[email protected], [email protected]
John R. Stauffer3
Infrared Processing and Analysis Center, California Institute of Technology, Mail Code
100-22, 770 South Wilson Avenue, Pasadena, CA 91125
[email protected]
and
Alison Sills
Department of Physics and Astronomy, McMaster University, 120 Main Street West,
Hamilton, Ontario L8S 4M1, Canada
[email protected]
ABSTRACT
1Based on observations obtained at the Anglo-Australian Telescope.
2Based on observations obtained at the Cerro Tololo-Interamerican Observatory, NOAO, which is operated
by the Associated Universities for Research in Astronomy (AURA), Inc., under cooperative agreement with
the National Science Foundation.
3Visiting Astronomer, Cerro Tololo Inter-American Observatory, NOAO.
4Present address, Physics and Astronomy Group, The Open University, Walton Hall, Milton Keynes MK7
6AA, United Kingdom.
-- 2 --
We report new measures of radial velocities and rotation rates (v sin i) for 51
F and early-G stars in the open cluster NGC 2516, and combine these with pre-
viously published data. From high signal-to-noise spectra of two stars, we show
that NGC 2516 has a relative iron abundance with respect to the Pleiades of
∆[Fe/H] = +0.04 ± 0.07 at the canonical reddening of E(B − V ) = 0.12, in con-
trast to previous photometric studies that placed the cluster 0.2 to 0.4 dex below
solar. We construct a color-magnitude diagram based on radial velocity mem-
bers, and explore the sensitivity of photometric determinations of the metallicity
and distance to assumed values of the reddening. For a metal abundance near
solar, the Hipparcos distance to NGC 2516 is probably underestimated. Finally,
we show that the distribution of rotation rates and X-ray emission does not differ
greatly from that of the Pleiades, when allowance is made for the somewhat older
age of NGC 2516.
Subject headings: open clusters and associations: individual (NGC 2516) -- stars:
distances -- stars: rotation -- Xrays: stars
1.
Introduction
Observations of stellar rotation rates in young clusters demonstrate that stars must lose
a considerable fraction of their initial angular momentum both before and after arrival on
the main sequence. In young clusters such as Alpha Persei (age ∼ 50 Myr) or the Pleiades
(age ∼ 100 Myr), stars at any mass are seen to rotate at a variety of rates (e.g., Stauffer,
Hartmann, & Jones 1989; Queloz et al. 1999; Terndrup et al. 2000), with most rotating slowly
(v sin i ≤ 10 − 15 km s−1), but a minority rotating much more quickly (v sin i ≥ 30 km s−1).
By the age of the Hyades (600 Myr), all but the lowest-mass stars have spun down to slow
rotation rates (Radick et al. 1987; Stauffer, Hartmann, & Latham 1987; Jones, Fischer, &
Stauffer 1996; Terndrup et al. 2000; Tinker, Pinsonneault, & Terndrup 2002). This picture is
reinforced by observations in clusters with ages between the Pleiades and the Hyades, such as
M 34 = NGC 1039 (Soderblom, Jones, & Fischer 2001) or NGC 6475 (James & Jeffries 1997),
although in these clusters the number of stars with measured rotation rates is smaller than
in the more well-studied systems. Here and throughout this paper, ages come from Meynet,
Mermilliod, & Maeder (1993), unless otherwise stated; we do not use ages derived from the
"lithium depletion boundary" technique (Stauffer, Shultz, & Kirkpatrick 1998; Barrado y
Navascu´es, Stauffer, & Patten 1999; Stauffer et al. 1999), because few clusters have ages
measured in this manner.
The current evolutionary paradigm for stellar angular momentum loss has a number
-- 3 --
of ingredients that are motivated by observations or theory. The main sequence spindown
is attributed to magnetic coupling between the stellar envelope and an ionized wind (e.g.,
Schatzman 1962; Weber & Davis 1967), which would cause a star to spin down over tens
to hundreds of millions of years. The existence of rapid rotators in clusters of age < 200
Myr requires that the angular momentum loss rate must be suppressed or saturated at high
rotation rates (Kawaler 1988; MacGregor & Brennan 1991). The presence of slow rotators
in young clusters points to an additional loss mechanism acting over lifetimes of a few Myr
(Cameron & Campbell 1993; Keppens, MacGregor, & Charbonneau 1995); motivated by the
theoretical work of Konigl (1991) and Shu et al. (1994), which showed that magnetic torques
can transfer angular momentum away from a protostar onto its accretion disk, this early loss
is termed "disk locking."
Much of the effort to understand the evolution of spin rates has concentrated on us-
ing clusters of various ages to set constraints on the duration of disk-locking and mass-
dependence of saturation used in the models (Krishnamurthi et al. 1997; Allain 1998; Sills,
Pinsonneault, & Terndrup 2000). An important but often unstated assumption in many of
these studies is that open clusters of different ages sample a single time sequence of angular
momentum loss. For this to be the case, the distribution of initial specific angular momenta
and disk locking lifetimes, as well as the properties of saturation and of winds, would be the
same in all open clusters (specifically, the dependence on stellar mass would be identical).
Testing this assumption has always been difficult since most of the available data were for
Alpha Persei, the Pleiades, and the Hyades, which mainly sample the loss mechanisms on the
main sequence. Consequently there have been a number of efforts underway to obtain new
observations of more clusters of similar age (e.g., Jones et al. 1997; James & Jeffries 1997), or
by extending observations to very young systems such as the Orion Nebula Cluster (Attridge
& Herbst 1992; Stassun et al. 1999; Herbst et al. 2000; Herbst, Bailer-Jones, & Mundt 2001;
Rebull 2001). Recently, Tinker, Pinsonneault, & Terndrup (2002) have shown that a single
set of parameters describing the angular momentum loss can explain the rotational evolution
from Orion to the Pleiades and Hyades.
The southern open cluster NGC 2516 provides an important additional laboratory for
exploring the evolution of angular momentum in young clusters. NGC 2516 is a fairly rich
cluster with a mass about twice that of the Pleiades, in which many possible members have
been identified photometrically (Jeffries, Thurston, & Hambly 2001). There have been radial
velocity studies of bright stars near the main-sequence turnoff (Gonz´alez & Lapasset 2000),
and a previous study of the rotation rates of F-G stars (Jeffries, James & Thurston 1998)
which we extend in this paper. The cluster has also been the target of several X-ray surveys
(Jeffries, Thurston, & Pye 1997; Micela et al. 2000; Harnden et al. 2001; Sciortino et al.
2001), and has a distance determination from Hipparcos parallaxes (Robichon et al. 1999).
-- 4 --
NGC 2516 has been of particular interest because several previous photometric studies
have suggested a low metallicity, a few tenths dex below solar (Cameron 1985; Jeffries,
Thurston, & Pye 1997; Jeffries, James & Thurston 1998; Pinsonneault et al. 1998).
In
contrast, a preliminary metallicity estimate from an equivalent width analysis of relatively
low signal-to noise spectra (Jeffries, James & Thurston 1998) indicated an abundance for
NGC 2516 that was solar within the errors. As in any other cluster, the cluster age from
the main-sequence turnoff depends sensitively on the metallicity. If NGC 2516 is near solar
metallicity, it has an age of 140 Myr, but would have an age about 180 Myr for [Fe/H]
≈ −0.3, a difference of 30%. Furthermore, accurate age determinations enter into many
other aspects of stellar evolution in open clusters, such as the relation between the initial
and final masses of white dwarfs and the maximum mass of white dwarf progenitors (Koester
& Reimers 1996; Jeffries 1997; von Hippel & Gilmore 2000; Weidemann 2000).
Determining the metallicity of NGC 2516 is also critical for using the cluster as a probe of
the metallicity dependence of angular momentum loss. Theory shows that, once on the main
sequence, the depth of the outer convective zone depends mainly on stellar mass, somewhat
on metallicity, and hardly at all on stellar age (e.g., Pinsonneault, DePoy, & Coffee 2001).
Stars of low metallicity have shallower convective zones at a given effective temperature,
and because low-metallicity stars of a given mass are also hotter and less luminous than
their solar-abundance counterparts, their convection zones should be thinner than higher-
abundance stars of the same color.
Here we present the results of a spectroscopic study of NGC 2516, which combines
parallel efforts led by two of us (DMT and RDJ). In § 2, we discuss the observations and
data reduction, and proceed in § 3 to derive a metallicity for NGC 2516 both from new
spectra and from analysis of the color-magnitude diagram. In § 4, we discuss the rotational
properties of the stars in our combined sample, and present our conclusions in § 5.
2. Observations and Data Reduction
2.1. Photometry
Photometry comes from a new catalog prepared by Jeffries, Thurston, & Hambly (2001)
from a CCD survey of 0.86 square degree of NGC 2516, obtained at the CTIO 0.9-m tele-
scope during three nights in January 1995. Conditions were photometric and the individual
CCD frames were flat-fielded with twilight exposures. Nightly corrections for extinction and
transformations to the Johnson V , B − V and Cousins V − IC systems were achieved using
many observations of standard fields from the Landolt (1992) compilation. The numbers of
-- 5 --
standards observed ensures that, for stars of the colors considered in this paper, statistical
errors in transforming to the standard colors and magnitudes can be neglected.
Aperture photometry was performed using small apertures (either 3 or 6 arcsec radius)
on a series of 25 mosaicked CCD fields, each covering an area of 13.5×13.5 arcminutes2 at a
scale of 0.4 arcsec/pixel. Comparison of stars in the overlapping regions of the CCD fields
gives us a good idea of the internal photometry errors. For the stars considered in this paper,
with V = 11 − 15, these internal errors are 0.022 mag in V , and 0.015 mag in B − V and
V − IC. Residuals to the linear transformation functions used to fit the Landolt standards
were less than 0.02 mag on all nights.
2.2. Spectroscopy at Cerro Tololo
Most of our new spectra were obtained in two runs at the Blanco 4 m telescope at Cerro
Tololo in March 1998 and January 1999. For both runs, we used the echelle spectrograph
with "red long" camera to produce spectra with an average dispersion of ≈ 0.08 A pixel−1.
The effective resolution as measured from the FWHM of arc lamp lines was 2.4 pixels, or
8.7 km s−1 at Hα. The spectra taken in 1998 used the G181 cross disperser (316 l mm−1),
and covered the wavelength interval 5840 -- 8150 A in 28 orders. In 1999, we used the 226-3
cross disperser (226 l mm−1) to obtain spectra from 5180 A to 8280 A in 39 orders. The
wavelength coverage in each order exceeded the free spectral range, so the wavelengths at the
ends of each order were sampled twice. In each run, we obtained spectra of bright field stars
with known radial velocity spanning the temperature range of our NGC 2516 targets, and
took spectra of the evening twilight sky. Most of the spectra from CTIO have signal-to-noise
of 15-25 per resolution element, but we obtained repeated long exposures of two stars for an
abundance analysis of NGC 2156, below.
Basic image processing and spectral reductions were performed with scripts written for
the IRAF5 and VISTA packages (Stover 1998). After removal of the overscan level and a
zero-exposure frame, each image was processed with a moving-box median filtering algorithm
to remove the many cosmic rays. The spectral orders were then traced using the exposures
of bright stars, and the sky and object apertures were defined for each object to maximize
the number of pixels contributing to the sky. Finally, the spectra were extracted using an
unweighted sum over the spectral aperture after background subtraction. This background
subtraction also removed scattered light from the spectra.
5IRAF is distributed by the National Optical Astronomy Observatory, operated by AURA, Inc., under
cooperative agreement with the National Science Foundation.
-- 6 --
The spectrograph at CTIO exhibits considerable flexure with telescope position, amount-
ing to ±1.5 pixels at the declination of NGC 2516 during each night. We corrected for this by
cross correlating selected orders of the extracted spectra containing the A and B atmospheric
bands, and then applied the measured pixel shifts to bring each spectrum to a uniform zero
point. The accuracy of this process was estimated by the scatter in the derived pixel shifts
for the four orders that had significant telluric absorption. The average error was typically
0.1 pixel, which is equivalent to 0.36 km s−1 at Hα, but was occasionally higher. This error
was later added in quadrature to the error in the radial velocity.
Wavelength calibration was achieved using ThAr lamp spectra in two steps. In the first,
lines were identified and a trial solution was obtained independently for each order. In the
second step, quadratic polynomials were fit to the central wavelength and dispersion as a
function of order number, and the line identifications and solutions for the wavelength at
each pixel were rederived using this fit as a starting point. In this step, as before, each order
had an independent wavelength solution.
Radial velocities were measured by cross correlation of seven extracted orders of each
spectra against the equivalent orders of bright, high signal-to-noise template stars in the lists
of Udry, Mayor, & Queloz (1999) and Stefanik, Latham, & Torres (1999). The seven orders
were chosen to have many stellar absorption features and to be relatively free of telluric
absorption, and had central wavelengths from 6040 A to 6778 A. Each order was separately
cross correlated and the results averaged. The scatter in the velocities for the separate orders
was used to estimate the error in the velocity. The lowest velocity errors, about 0.5 km s−1,
were found for the most narrow-lined stars (i.e., the most slowly rotating ones) and/or those
stars with higher signal to noise. For stars of moderate to rapid rotation, the radial velocity
error is typically about 1/20 the value of v sin i. The total error was taken as the quadrature
sum of the radial velocity error and the error in the flexure correction. The average error for
the entire sample was about 1.1 km s−1.
The zero point of the radial velocity system was defined by the stars HR 2593 (spectral
type K2), HR 4540 (F8), HR 4695 (K1), and HR 5353 (G5). The estimated external error
for the velocities in this system is 0.4 km s−1.
The width of the cross correlation between each star and the template was used to deter-
mine the rotational velocity. The calibration of the relation between the correlation widths
and v sin i was determined by convolving the spectrum of another high signal-to-noise spec-
trum of a slowly-rotating star with rotational broadening functions, then cross-correlating
those artificially broadened spectra with the template star spectrum (cf. Hartmann et al.
1986). The broadening functions are given by Gray (1992); we adopted a limb-darkening
coefficient ǫ = 0.6.
-- 7 --
2.3. Spectroscopy at the AAT
Data were taken on the AAT on 2000 February 23-24. The UCL Echelle Spectrograph
with 79 lines mm−1 grating was used at the Coud´e focus, with a MITLL3 chip. The central
wavelength was 6150A in order 37. A 1.2′′ slit was used, giving a resolution of 0.13A at
0.036A pixel−1 (in order 33).
The data were reduced at the Keele Starlink node, using the echomop, figaro and
uclsyn packages (Mills, Webb & Clayton 1997; Smith 1992). As with the CTIO data,
the spectra reduction here included background subtraction (including scattered light), and
wavelength calibration.
Radial velocities were measured using cross-correlation techniques in the spectral range
5210-5340A and 5460-5600A, which contain many neutral metal lines. The standard tem-
plates used to determine heliocentric RVs were radial velocity standards HR 1829 (G5 V)
and HR 4540 (F9 V). Internal errors are dominated by small shifts in the wavelength calibra-
tions during exposures amounting to about 1 km s−1. External errors, found by comparing
standard star observations, are estimated to be 0.4 km s−1.
Rotational velocities were obtained by fitting a synthetic spectrum to the data, allowing
the line abundances and rotational velocity to vary until the minimum χ2 value was obtained.
The atmospheres used were those of Kurucz (1993) and used a mixing length approximation
with no overshooting. Oscillator strengths were fixed by comparison with a solar spectrum
taken with the same instrumentation.
The spectral range used was 5260-5290A, which was in the order with the deepest
lines. Temperatures for the synthesis were calculated from the B − V photometry given
in the catalog of Jeffries, Thurston, & Hambly (2001), using the prescription of Saxner &
Hammarback (1985), which includes a dependence on [Fe/H]. For this analysis, we assumed
[Fe/H] = −0.32, but the results would be nearly the same had we assumed solar abundance.6
Errors in v sin i were obtained by finding the v sin i at which the χ2 values increased by an
amount appropriate for the number of degrees of freedom. This was added in quadrature
to an uncertainty of around 0.7 km s−1, which arises from uncertainties in the effective
temperature put into the synthesis (a change in B − V of 0.02 changes the temperature by
75 K at 6500 K).
6The metallicity of the atmosphere used in the analysis has little effect on the line widths, but would
change the relation between color and effective temperature. Adopting solar metallicity would have made
the models ≈ 100 K hotter. The difference in the intrinsic width of the lines for this temperature change
does not produce a significant effect in the derived value of v sin i.
-- 8 --
2.4. Results and internal checks
Table 1 presents our measurements in NGC 2516. Column 1 lists the name of each star,
where DK indicates the list of Dachs & Kabus (1989), E indicates Eggen (1972), and the
remaining names are from the photometric catalog of Jeffries, Thurston, & Hambly (2001)
or those used in Jeffries, James & Thurston (1998). Columns 2 -- 6 display the source for the
data, the radial velocity and its error, and the rotational velocity and its error. Finally, the
last five columns show the photometry, catalog entry, and a photometric membership flag for
each star from Jeffries, Thurston, & Hambly (2001). The first part of the table shows stars
which we class as cluster members because their radial velocities are statistically consistent
with the cluster mean velocity, while the second part shows stars with discrepant radial
velocities. These memberships were assigned by an iterative scheme that determined which
stars were within 3σ of the cluster mean velocity. We determined the mean by combining
our observations here with the velocities for members in Jeffries, James & Thurston (1998).
The reported errors were increased by adding 1 km s−1 in quadrature to account for the
likely velocity dispersion in the cluster (Stauffer et al. 1997).
Six of the stars in the CTIO sample had previously been observed by Jeffries, James
& Thurston (1998). The weighted mean difference in radial velocity in the sense (CTIO --
previous) is +1.1 kms, indicating that the CTIO data have a slightly higher zero point in
velocity than the earlier work. The weighted scatter in the difference is 0.9 km s−1, where
a value of 1.3 km s−1 would be the expected error from the random errors in the velocities.
Table 2 lists the mean cluster velocity from this study and earlier works; the error in that
table includes only random errors. The mean velocity for our full sample is slightly higher
than derived in Jeffries, James & Thurston (1998), but is not statistically significant given
the external errors here and that paper (≈ 0.4 km s−1). Our cluster velocity is distinctly
greater than the value of +19 km s−1 quoted in the Lynga (1987) catalog, and is somewhat
greater than the value of 22.0 ± 0.2 km s−1 derived for early-type stars and red giants in
NGC 2516 by Gonz´alez & Lapasset (2000). The latter study included a correction of −0.6
km s−1 for the gravitational redshift of main-sequence stars, a correction we do not apply
here. Had we applied such a correction, our mean cluster velocity would still be slightly
higher than in Gonz´alez & Lapasset (2000).
Figure 1 shows the correlation of the new and older values of v sin i; the solid line on that
figure indicates equality and is not a fit to the data points. Five of the stars agree within the
errors, while one star (DK 206) has a v sin i value from CTIO that is considerably greater
than found previously. We are unable to discern the source of this discrepancy, although we
have verified that the recorded coordinates for this star on the CTIO run match the catalog
position of DK 206.
-- 9 --
All but two of the stars we observed were listed as probable members in Jeffries,
Thurston, & Hambly (2001), where membership was determined by proximity to the main
sequence in both the V , B − V and V , V − I color-magnitude diagrams and in the B − V ,
V − I two-color plot. The exceptions are DK 573 and DK 742, which passed two of the tests.
Of the stars listed as non-members (lower part of Table 1), most are photometric mem-
bers in Jeffries, Thurston, & Hambly (2001). Three of these (DK 417, DK 669, and DK 919)
have velocities within 20 km s−1 of the cluster mean. Of these, only DK 417 has a de-
tected rotational velocity (v sin i = 17.9 km s−1), and so may be a binary member of
NGC 2516. DK 669 is listed as a photometric binary candidate in Jeffries, Thurston, &
Hambly (2001). Several other stars have more discrepant velocities but also exhibit signifi-
cant rotation (DK 306, DK 503, and DK 508); these may be candidate short-period binaries.
Of these three, only DK 508 is a binary candidate from photometry. Second-epoch radial
velocities would be helpful to further elucidate their nature. To be conservative, we have not
included the non-member stars in the discussion below.
3. Analysis
3.1. The abundance of NGC 2516
We obtained repeated, long exposures of two slowly-rotating stars (DK 320 and DK 325);
the resulting average spectra are of sufficient quality (peak S/N of about 70 per pixel) to
attempt a fine spectroscopic abundance analysis. From the point of view of open cluster
studies, the best approach is to define the abundance relative to other well studied clusters
such as the Pleiades. To do this, we performed a differential iron abundance analysis with
respect to the Sun, using lines for which equivalent widths have been published for Pleiades
stars. We analyzed our measured widths in NGC 2516 and those in the Pleiades using the
same model atmospheres and adopted temperature scale.
We chose to work with a set of six isolated, unblended Fe I lines concentrated around
6700 A (specifically 6677 A, 6703 A, 6705 A, 6726 A, 6750 A, and 6752 A). Equivalent
widths for these lines in a set of Pleiades F dwarfs are given by Boesgaard, Budge, & Burck
(1988) and by Boesgaard, Budge, & Ramsay (1988). We found that many of the other lines
employed by Boesgaard & Friel, specifically those with λ > 7000 A, were blends, and that
our measurements of line strengths in twilight solar spectra were not in agreement with the
values quoted by Boesgaard & Friel.
We measured the equivalent widths of our chosen lines in both of the two targets and
in a twilight solar spectrum using the same instrumentation and reduced in a similar way.
-- 10 --
We compared the solar equivalent widths with those from the Kitt Peak Solar Atlas, finding
good agreement in all cases. We then used the one-dimensional homogeneous LTE model
ATLAS9 model atmospheres (Kurucz 1993) to calculate the iron abundances. The atmo-
spheres incorporated the mixing length treatment of convection with α = 1.25. We did this
in a differential way with respect to the Sun, assuming solar parameters of Teff = 5777 K,
log g = 4.44, and a mictoturbulent velocity of 1.0km s−1. The gf factors for the lines were
tuned to give a solar iron abundance of 7.54 (Milford, O'Mara, & Ross 1994) on the usual
logarithmic scale where the abundance of hydrogen is 12.
The equivalent widths for our target stars and for the Pleiades F stars were then fed
into the same models with Teff determined from B −V via the Saxner & Hammarback (1985)
relationship, assuming a solar metal abundance and log g = 4.5. We assumed E(B − V ) =
0.12 for NGC 2516 (Lynga & Wramdemark 1984; Dachs & Kabus 1989) and E(B −V ) = 0.04
for the Pleiades (Crawford & Perry 1976; Stauffer & Hartmann 1987) to obtain intrinsic
colors. We allowed the microturbulence to be a fitting parameter, requiring that the derived
iron abundance was independent of line equivalent width. We found that in all cases the
microturbulence was in the range 1.0 to 1.75 km s−1 with an uncertainty of about 0.25 km
s−1. To investigate systematic errors, we re-determined abundances using different values
of Teff , log g, and microturbulence. A temperature variation of ±100 K, corresponding to a
change in intrinsic color of ∓0.03 mag, led to ∆[Fe/H] = ±0.07; a change in log g = ±0.2 gave
∆[Fe/H] = ±0.03; a change in the microturbulence of +0.25 km s−1 led to ∆[Fe/H] = ∓0.04.
We find [Fe/H] = −0.02 ± 0.03 for DK 320 and [Fe/H] = +0.13 ± 0.04 for DK 325,
where the error is the standard error in the mean from the six lines. To these errors we
add about 0.08 dex to account for errors in the photometry, gravity, and microturbulence.
The mean for NGC 2516 from the two stars is therefore [Fe/H] = +0.05 ± 0.06. To put
this onto an absolute abundance scale we should then also have to consider errors in the
assumed temperature scale, atmospheric models, and assumed mean reddening. We should
also consider the consequences of using [Fe/H] = 0 to calculate the temperatures from the
intrinsic colors. This is in fact a small correction, because each 0.1 dex decrease in metallicity
only decreases the assumed effective temperature by ≈ 30 K. Thus if we had assumed initially
that [Fe/H] = −0.3 for the purpose of calculating Teff, we would have derived [Fe/H] = −0.01
from the spectra, which is clearly inconsistent with the original assumption.
There were six Pleiades stars for which all six Fe I lines had equivalent width mea-
surements. We find a mean [Fe/H] = 0.01 ± 0.03, assuming intrinsic color errors of ±0.02
and assuming similar gravity and microturbulence errors to the NGC 2516 stars. This is in
reasonable agreement with [Fe/H] = −0.034 ± 0.024 found by Boesgaard & Friel (1990), who
also employed the Saxner & Hammarback (1985) temperature scale in their analysis.
-- 11 --
For the purposes of comparison, the ratio of iron abundance in NGC 2516 and the
Pleiades does not depend on the adopted temperature scale nor on the model atmospheres
employed, because both sets of data were treated in the same way.7 The logarithmic ratio of
the iron abundances in NGC 2516 to the Pleiades is therefore ∆[Fe/H] = +0.04±0.07, and so
we therefore conclude that the metallicity of NGC 2516 is roughly solar abundance. For our
final value, we combine this differential measurement of [Fe/H] and the Boesgaard & Friel
scale for the Pleiades and conclude that the metallicity of NGC 2516 is [Fe/H] = +0.01±0.07.
The only additional error is about ±0.02 in [Fe/H] for every ±0.01 change in the relative
reddening values for each cluster. We discuss the reddening in further detail in the next
section when we look at the CMD for NGC 2516, but for now we note that if we took an
extreme reddening for our two stars of E(B − V ) = 0.19 (Schlegel, Finkbeiner, & Davis
1998), we would derive ∆[Fe/H] = +0.15 with respect to the Pleiades. We adopt an error
in the exction of ±0.02 so the error in our spectroscopic determination of [Fe/H] rises to
±0.08. Nevertheless we caution the reader that this result has been derived from only two
stars, and there is a great need to increase this sample. Unfortunately, most of the F stars
have spectra which have large rotational broadening, so it may be necessary to pursue this
work on the fainter G stars with larger telescopes.
Our new abundance is consistent with at least some other determinations of the abun-
dance of NGC 2516. Twarog, Ashman, & Anthony-Twarog (1997) give [Fe/H] = +0.060 ±
0.030 from DDO photometry of two stars. Nissen (1988) reports uvby − β photometry of 6
stars in NGC 2516, deriving [Fe/H] = +0.06 ± 0.06 and E(b − y) = 0.081 ± 0.016, which
corresponds to E(B − V ) = 0.11 ± 0.02. The Nissen et al. result, however, disagrees with
that of Lynga & Wramdemark (1984), who also report uvby − β photometry in NGC 2516
and derive [Fe/H] = −0.28 using about the same extinction as Nissen (1988).
3.2. The distance and reddening to NGC 2516 and its photometric metallicity
We now present a new determination of the cluster distance and photometric metallicity
using an updated version of the main-sequence fitting method outlined in Pinsonneault et
7For example, an alternative calibration of the color-temperature relation is provided by Houdashelt, Bell,
& Sweigart (2000), who provide polynomial fitting functions in both B − V and V − I. For the two stars
in question, this temperature scale is 66 K cooler in B − V that that from Saxner & Hammarback (1985)
used in the abundance analysis. For the entire sample the mean difference is ≤ 25 K for any reddening
E(B − V ) ≤ 0.2;
it just so happens that our two stars with high S/N spectra have colors where the
temperature difference between the two B − V scales is a maximum.
-- 12 --
al. (1998). The full details of the method may be found in Pinsonneault et al. (2002), and
a preliminary determination for NGC 2516 is discussed in Pinsonneault, Terndrup & Yuan
(2000).
The process of isochrone fitting separately in B − V and V − I yields a photometric
metallicity for a cluster since the luminosity of the main sequence at fixed color changes more
rapidly in B − V (where ∆MV /∆[Fe/H] ≈ 1.3) than it does in V − I (∆MV /∆[Fe/H] ≈ 0.6).
The signature of a metal poor cluster would then be a distance modulus derived in B − V
that was greater than found in V − I. Indeed, Jeffries, James & Thurston (1998) used this
technique to derive [Fe/H] = −0.18 ± 0.05 for NGC 2516, using the smaller sample of radial
velocity members available then, while Pinsonneault, Terndrup & Yuan (2000) found [Fe/H]
= −0.26 using the same data; both studies used E(B − V ) = 0.11. As these are both at
variance with our spectroscopic abundance (above), we wish to explore the distance and
metallicity determinations in some detail.
Our approach here is similar in approach to that employed by Jeffries, Thurston, &
Hambly (2001) in the determination of candidate photometric members of NGC 2516. In that
study, they generated empirical isochrones starting with the Siess, Dufour, & Forestini (2000)
models, and then calibrated the isochrones with multicolor photometry in the Pleiades,
assuming a Pleiades distance modulus of 5.6. They derived a distance to NGC 2516 by
fitting a 150 Myr isochrone to the main-sequence appearing in B − V and V − I CMDs,
which is relatively distinct from the background except at faint magnitudes. The result of this
calculation determined the relative distance between the Pleiades and NGC 2516. The result
depended on the metallicity: for solar abundance, the distance is 8.10 ± 0.05 in both B − V
and V − I, while for half solar the distances in B − V and V − I were respectively 7.85 ± 0.05
and 7.90 ± 0.05. Both calculations assumed E(B − V ) = 0.12, and a particular choice of the
extinction law. By comparison, the Hipparcos parallax translates to a distance modulus of
(m − M)0 = 7.70 ± 0.16 (Robichon et al. 1999), which is close to the value of 7.77 ± 0.11
found by Sung et al. (2002). This last study presented UBV I photometry for NGC 2516
independently of Jeffries, Thurston, & Hambly (2001), and derived a photometric metal
abundance of [Fe/H] = −0.10 ± 0.04 by comparison of the main sequence to uncalibrated
isochrones from the Padua group. We will have more to say about the Sung et al. (2002)
analysis below.
We begin the discussion by comparing the photometry of our radial velocity members
to isochrones at the distance measured by Hipparcos. Figure 2 shows the color-magnitude
diagram in V0 and (B −V )0 for the combined sample, where the filled points display photom-
etry for our radial-velocity members and the open points show the Dachs & Kabus (1989)
photometry for main-sequence members from the radial velocity study of Gonz´alez & La-
-- 13 --
passet (2000). The stars deemed non-members from our radial velocities are displayed as
plusses. Figure 3 displays a CMD in V0, (V − I)0 for the magnitude range containing our
sample only. The photometry has been dereddened according the prescription of Bessell,
Castelli, & Plez (1998, their Appendix F) using a value E(B − V )0 = 0.12 for hot stars
and including the dependence of extinction and reddening on stellar color. In B − V , there
is a tight single-star sequence with a distinctly brighter binary sequence; in V − I there
is no such clear separation and the main-sequence has a fairly wide range of colors. The
dashed lines on Figures 2 and 3 show an isochrone for [Fe/H] = 0.0, calibrated as described
below, and relocated to the Hipparcos distance. We conclude from Figures 2 and 3 that the
Hipparcos distance and the similar Sung et al. (2002) value are correct only if NGC 2516
is metal poor (about [Fe/H] = −0.3), provided that the distance modulus to the Pleiades
is near 5.6. Alternatively, NGC 2516 may share the Pleiades "problem" (Pinsonneault et
al. 1998) of having a Hipparcos parallax that makes the cluster considerably fainter than
expected for its metal abundance as derived in this paper.
The isochrones were derived from evolutionary tracks incorporating angular momentum
transport and an updated equation of state (Sills, Pinsonneault, & Terndrup 2000; Tinker,
Pinsonneault, & Terndrup 2002). The tracks are first used to generate isochrones in the
theoretical plane (i.e., MV , Teff). An initial calibration to BV RI colors comes from the Leje-
une, Cuisinier, & Buser (1997) model atmospheres. The colors are then further corrected to
match the photometry of single star members of the Hyades8 cluster; this step is necessary
because the isochrone shapes after initial calibration still do not match the observed cluster
main sequence at all luminosities (e.g., Terndrup et al. 2000, their Figs. 4 and 5). Parallaxes
for individual Hyades members, which have typical errors of 2%, were used to correct the
photometry to the distance of the dynamical center of the cluster which has a distance mod-
ulus 3.34 ± 0.01, or d = 46.6 ± 0.2 pc (Perryman et al. 1998). The empirical corrections were
generated for a theoretical isochrone with [Fe/H] = +0.13 (Boesgaard & Friel 1990), with
scaled solar abundances and the solar helium abundance. The empirical color corrections
can then be applied to isochrones of any age and metallicity. For metallicities within a few
tenths dex of solar the empirical corrections to the colors are assumed to be independent of
metallicity; this is equivalent to assuming that the effect of line blanketing on the colors is
correct in the model atmospheres employed in the initial calibration of the isochrones.
In the Pinsonneault et al. (2002) method, cluster distances are then found by determin-
8In its full detail, the Pinsonneault et al. (2002) method also includes fainter Pleiades stars to extend
the empirical calibration to redder colors than are available for Hyades stars with Hipparcos parallaxes; this
consideration does not, however, make a difference here to NGC 2516 since our survey only extends to about
(B − V )0 = 0.9, a color range completely spanned by Hyades stars with excellent Hipparcos parallaxes.
-- 14 --
ing V0 − MV for each star, where MV is the absolute magnitude of the empirically calibrated
isochrone at the dereddened color of the star, and V0 is the dereddened magnitude. The
cluster distance is found separately in B − V and V − I from the median value of V − MV
in color bins. Stars that are more than 0.1 mag from the median are rejected, which reduces
the effect of binaries on the distance determination. The distance modulus is derived for
isochrones of several different metallicities. The cluster photometric metallicity is taken as
the metallicity of the isochrone for which the distance moduli determined from B − V and
V − I agree. In addition, it is generally possible to use the narrowness of the main sequence
in B − V as a measure of whether the shape of a particular isochrone matches the main
sequence data in that color; a shape mismatch is indicated by finding that hV0 − MV i is a
function of color.
There are two problems which arise in determining the distance and photometric metal-
licity of NGC 2516. First, the cluster has a markedly higher reddening than most of the
clusters in the original study of Pinsonneault et al. (1998). The derived photometric metal-
licity depends sensitively on the adopted reddening value, the ratio of total to selective
extinction in the various photometric bands, and whether there is differential reddening
towards the cluster. Second, the large span in color of the main sequence in V − I (Fig.
3) means that there will be a wide range of statistically permitted values of the apparent
distance modulus in the (V , V − I) CMD; as a consequence, the photometric metallicity
will have a relatively large error compared to clusters with narrow sequences. The width of
the main sequence in V − I may be caused by a significant binary fraction in the cluster,
any observational selection favoring binaries over single stars, or by significant differential
extinction towards NGC 2516.
Previous studies of the extinction towards NGC 2516 have generally arrived at values
near E(B−V ) = 0.11 or 0.12 (Dachs & Kabus 1989; Sung et al. 2002, and references therein).
In contrast, the Schlegel, Finkbeiner, & Davis (1998) dust-emission maps in the direction
of NGC 2516 indicate a higher average reddening hE(B − V )i = 0.21 to our spectroscopic
members, with a variation of ∆E(B − V ) = 0.03 towards our target stars. That the average
reddening is higher than that obtained from the hot turnoff stars in the cluster is not,
perhaps, a surprise, since the maps are based on the integrated infrared emission out to
infinity and the cluster is located at a distance of about 400 pc at a galactic latitude of
b = −16◦. Also, some studies (Arce & Goodman 1999; von Braun & Mateo 2001) have
claimed that the Schlegel, Finkbeiner, & Davis (1998) maps overestimate the extinction by
30 -- 50% when AV > 0.5 (equivalently when E(B − V ) ≥ 0.15).
In this analysis, we attempted to fit isochrones of a wide variety of metallicities using
the sample of radial velocity members in Table 1, and allowed the reddening to vary from the
-- 15 --
previously determined values. Low metallicities ([Fe/H] ≤ −0.4) provided a poor fit to the
isochrone shapes at any value of the reddening. Statistically acceptable fits were obtained
for metallicities between [Fe/H] = −0.3 and +0.1 at reddenings between E(B − V ) = 0.12
and 0.18. For the canonical reddening of E(B − V ) = 0.12, the photometric metallicity for
NGC 2516 is [Fe/H] = −0.05 at a distance of V0 − MV = 7.93. We illustrate this solution in
Figures 2 and 3, where the best fitting isochrone is displayed as a solid lines in each figure.
In B − V , this isochrone runs through the single-star sequence, and runs along the lower
edge of the main sequence in V − I. The match of the isochrone shapes to the photometry
is shown in Figure 4, which shows the individual values in V0 − MV in each color. There is
no significant trend with magnitude in the residuals in either color.
We extensively explored the sensitivity of the determination of the distance and photo-
metric metallicity to choices of reddening and the reddening law. The results are summarized
in Table 3. The smallest source of error is the accuracy of finding the median apparent dis-
tance modulus. The largest source of error is the value of the extinction itself. Adopting
higher extinctions increases the photometric metallicity appreciably but reduces the derived
distance only slightly; the effect on the distance modulus is smaller since the slope of red-
dening vector is only somewhat different from the average slope of the main sequence. An
equally important source of error is the steepness of the adopted extinction law, which may
be characterized by the value of RVI ≡ E(V − I)/E(B − V ).
Increasing RVI results in
a smaller distance and a significantly higher photometric metallicity. The final values we
derive for NGC 2516 are [Fe/H] = −0.05 ± 0.14 and V0 − MV = 7.93 ∓ 0.14.
Adopting an E(B − V ) error of ±0.02, then the photometric metallicity we found for
NGC 2516 is consistent, within the errors, with the value of +0.01 ± 0.08 we find from our
spectroscopic analysis. The main reason we derive a higher metallicity from the CMD here
than did Pinsonneault, Terndrup & Yuan (2000) is because we employ the Bessell, Castelli,
& Plez (1998) extinction law, for which RVI = 1.34 at the mean V − I of our sample, in
contrast to the value RVI = 1.25 used in the previous determinations of the metallicity from
the CMD. In our analysis, we took ±0.04 for the error in RVI, representing the difference
between the adopted value and the Dean, Warren, & Cousins (1978) value of RVI = 1.30 for
the average color in our sample. The reddening law we used here has A(V )/E(V −I) = 2.47,
on average, which is equal to the value of 2.49±0.02 determined empirically from observations
of stars in the bulge by Stanek (1996).
The distance we determine is 1.5σ higher than the Hipparcos value, which is not a
statistically significant difference by itself. This difference corresponds to a difference in
parallax of 0.29 mas, quite a bit smaller than the 1−2 mas systematic errors in the parallaxes
needed to explain the difference between the Hipparcos and isochrone-fitting distance to the
-- 16 --
Pleiades (Pinsonneault et al. 1998; Narayanan & Gould 1999). Reconciling the photometric
distance with the Hipparcos distance would require the photometric metallicity to be lower
than we determine, which would make that metallicity inconsistent with our spectroscopic
value.
If instead we adopted our spectroscopic value and fit a solar abundance isochrone to the
narrow B − V sequence in Figure 2, we would obtain a more precise distance of (V0 − MV ) =
8.05 ± 0.11, which is 2.5σ higher than the Hipparcos distance.
Recently, Sung et al. (2002) analyzed new UBV I photometry of NGC 2516, and used
the Padua group isochrones to derive a photometric metallicity of [Fe/H] = −0.10 ± 0.04
and a distance modulus V0 − MV = 7.77 ± 0.11. Therefore despite finding a relatively
high abundance for NGC 2516, their distance is consistent with the Hipparcos value.
In
our opinion, their analysis is seriously flawed, even though their distance determination is
marginally equal to ours within the errors. Here we have demonstrated that one can achieve
an excellent fit to the CMD in both B −V and V −I from isochrones calibrated on the Hyades
and the Pleiades. An inspection of Figure 2 in Sung et al. (2002) shows that the isochrones
they adopted do not simultaneously fit the main sequences in B − V and V − I and are a
poor match to the shape of the main sequence in B −V or U −B. Their determination of the
distance rests primarily on fitting an isochrone to the (V, V − I) CMD; the corresponding
isochrone in B − V is significantly too bright (or red) and misses the bulk of the main
sequence over at least 5 mag in V . They then attribute the difference between the isochrone
and the stellar locus to a UV excess, and claim that this is a general feature of young stellar
clusters. Finally, they determine the cluster metallicity by rejecting most of the stars in the
cluster as having a UV excess, which necessarily means that they derived a value near solar
because their sample is contaminated with nonmembers. In a separate paper (Stauffer et
al. 2002), we present evidence for a blue excess in young, low-mass stars, mainly from data
for late-K dwarfs in the Pleiades. We find no evidence for blue excesses in G and early K
dwarfs, and hence we do not believe that this phenomenon should affect the results we have
presented for NGC 2516.
3.3. Other metallicity estimates
Debernardi & North (2001) have recently published a detailed study of the eclipsing
binary system V392 Car, which is a member of NGC 2516. They compared their derived
luminosity for the members of this system against stellar structure models using a distance
close to the Hipparcos measurement, and derived a metallicity of [Fe/H] = 0.0 ± 0.1. We
performed a similar analysis, instead comparing the stars' measured radii to that in the
-- 17 --
isochrones we used for the distance analysis. The result is [Fe/H] = −0.05 ± 0.05 for these
stars by the constraint that their radii match the isochrone radii, consistent within the errors
with our spectroscopic determination.
4. Stellar rotation in NGC 2516
Following the approach of Soderblom, Jones, & Fischer (2001) and Solderblom et al.
(1993), we display in Figure 5 the correlation between v sin i and color for NGC 2516 and
other well-studied open clusters. From top to bottom are shown the Pleiades, NGC 2516,
M 34 and the Hyades (600 Myr). The data sources are as follows: Pleiades: Queloz et
al. (1999), supplemented with new observations in Terndrup et al. (2000); NGC 2516: this
paper; M 34: Soderblom, Jones, & Fischer (2001); Hyades: v sin i data from Kraft (1965,
1967), and rotational periods from Radick et al. (1987) converted to v sin i. The solid line
in each panel is a representation of the Hyades data, while the dashed line indicates the
location of the rapid rotators in the Pleiades. For NGC 2516, we adopted E(B − V ) = 0.12
(above).
In Fig. 5, the time evolution of angular momentum from the Pleiades to the Hyades
(100 to 600 Myr) manifests itself as a gradual reduction in v sin i for the slow rotators, and
a significant decline in the numbers of rapid rotators. NGC 2516, like the Pleiades, has
a number of rapid rotators, and these have rotation speeds about 70% of the stars in the
Pleiades with similar colors. The number of rapid rotators declines markedly by the age of
M 34, which has an age estimated from 180 Myr (Meynet, Mermilliod, & Maeder 1993) to
∼ 250 Myr (Jones & Prosser 1996). In M 34, the upper bound of rotation rates is about a
factor of two lower than it is in the Pleiades. Our observations in NGC 2516 do not extend
to the faint magnitudes at which the K dwarfs in the Pleiades and M 34 are still rapidly
rotating.
In Figure 6, we compare the rotation rates of stars in the Pleiades, NGC 2516, M 34
and the Hyades with theoretical models. The data sources are the same as in Figure 5.
The conversion between color and effective temperature is taken from Houdashelt, Bell, &
Sweigart (2000). The models are taken from Sills, Pinsonneault, & Terndrup (2000), and
are for solar metallicity stars. The adopted ages of the clusters are shown in figure 6, where
we have taken the Jones & Prosser (1996) age of 250 Myr for M 34 as in their paper. In the
models the stars are assumed to be differentially rotating. The loss of angular momentum
from the surface was modelled using a saturated magnetic wind loss law (Chaboyer, Demar-
que, & Pinsonneault 1995) with a mass-dependent saturation threshold ωcrit which follows a
Rossby scaling down to M = 0.5M⊙ and then chosen to match the low mass rapid rotators
-- 18 --
in the Hyades. We used five different values of the disk-locking lifetime, during which the
star is assumed to be rotating at the same rate as its proto-stellar disk. The lifetimes are
chosen to span the expected range of proto-stellar disk lifetimes, up to 10 Myr.
The overall impression from Figure 6 is that the rotation rates of the stars in NGC 2516
are similar to what would be expected at an age near 140 Myr if the angular momentum
evolution of intermediate mass stars (M = 0.6 − 1.1M⊙) in all clusters was similar. In this
plot, M 34 has a few stars with rotation rates significantly higher than the fastest-spinning
model. This may represent a departure from the suggested uniformity of the Pleiades :
NGC 2516 : Hyades evolution, or else the age of M 34 is closer to the value of 180 Myr
from Meynet, Mermilliod, & Maeder (1993) than the 250 Myr shown here. Establishing that
different star clusters represent time snapshots of a universal evolution of angular momentum
thus would require better age estimates for these clusters than is currently available.
The data for both NGC 2516 and M 34 probe only the intermediate mass stars, and
do not extend to the low mass stars where the standard Rossby scaling breaks down. There
is a substantial change in behavior of the low mass models in this age range depending on
the choice of the saturation threshold (standard Rossby scaling vs. the Sills, Pinsonneault,
& Terndrup (2000) fit to the Hyades data).
If initial conditions for star formation were universal, we would expect that the distri-
bution of disk lifetimes would be the same from cluster to cluster, and should not depend
on the age of the cluster. It appears that the Pleiades and NGC 2516 have similar fractions
of stars with disk lifetimes between 0 and 1 Myr. However, the expected rotation rates
for stars with disk lifetimes longer than 1 Myr are slow enough that they overlap with our
observational limits. We cannot say whether the distribution is the same without better
resolution of rotation rates or photometric period determinations.
In Figure 7 we plot the fractional luminosity in X-rays (LX/Lbol) for NGC 2516 against
the inverse of Rossby number, which is defined as the ratio of the rotation period to the
turnover of time of a convective cell at the base of the convective zone. The X-ray luminosities
were taken from Jeffries, Thurston, & Pye (1997); filled points show measured values, while
filled triangles show upper limits on the X-ray detection. The open points show equivalent
data for a set of Pleiades stars with known rotation periods (Krishnamurthi et al. 1998). In
order to treat both data sets the same way, we ignored the rotation periods for the Pleiades
stars and instead began with v sin i values for these. The measured values for v sin i were
used to compute a rotation period using the empirically calibrated isochrones employed
above, and the isochrones were used to generate relations between effective temperature
and the B − V and V − I colors. The convective overturn time was computed from the
effective temperature using the prescription in Kim & Demarque (1996). We did not apply
-- 19 --
a statistical correction for projection to the data, which would have increased the v sin i
values by a factor of 4/π = 1.27. The conclusion from Figure 7 is that both the Pleiades
and NGC 2516 have the same relation between convection zone depth and X-ray activity.
The Pleiades stars plotted here have a higher average rotation rate and X-ray luminosity,
but the (fewer) rapid rotators in NGC 2516 are similar. The bulk of the NGC 2516 data lie
amongst the more slowly-rotating Pleiades; again there is no obvious difference between the
two clusters.
5. Summary and discussion
The principal result of this paper is the demonstration that NGC 2516 has approximately
solar metallicity, which we showed from an equivalent width analysis on high signal-to-noise
spectra and from the location of the main sequence on CMDs. While the metallicity from
the latter method has rather large errors, both random and systematic, it is nevertheless
consistent with the metallicity found from our spectra. We also showed that the rotational
and X-ray properties of the F and early G stars in NGC 2516 are consistent with what would
be expected at an age near 140 Myr if angular momentum evolution in different clusters is
a uniform time sequence.
Why, then, did previous studies come up with values of the metallicity that were almost
always metal poor? Both Cameron (1985) and Jeffries, Thurston, & Pye (1997) concluded
the cluster was metal poor by modeling the colors in the U − B, B − V plane, where the
data were a combination of photographic and photoelectric photometry. The F stars in
NGC 2516 clearly showed an ultraviolet excess compared to similar stars in the Pleiades
(Jeffries, Thurston, & Pye (1997), their Figs. 10 and 11). The new CCD photometry that
we report here from Jeffries, Thurston, & Hambly (2001) compares quite favorably to the
previous values in B − V over the magnitudes of stars in our sample, though the new study
does not include photometry in U for an independent check of the zero point in that band.
On the other hand, our conclusion that NGC 2516 has approximately solar metallicity
helps explain many other observations of the cluster. For example, Micela et al. (2000)
analyzed Rosat HRI observations of NGC 2516 stars, and showed that the X-ray activity in
this cluster is consistent with that observed in dK and dM stars of similar age, contrary to
the expectation that the convection properties in NGC 2516 stars would be different than
in solar-metallicity clusters. The depletion patterns of Li in this cluster are consistent with
a metallicity near solar (Jeffries, James & Thurston 1998), but would require nonstandard
mixing scenarios if the cluster were metal poor.
-- 20 --
Extending our work to lower masses where rapid rotation is observed (e.g., Figure 5)
is clearly warranted. Since NGC 2516 is significantly more distant than the Pleiades, v sin i
studies amongst the M dwarfs would require telescopes in the 8-10m range (M dwarfs have
V ≥ 14 in the Pleiades, or V > 16.6 in NGC 2516).
We would like to thank the directors and staff of both CTIO and the Anglo-Australian
Observatory. AF was a Particle Physics and Astronomy Research Council (PPARC) research
student during this work. DMT and MHP acknowledge partial support from the National
Science Foundation, via grants AST-9528227 and AST-9731621 to The Ohio State University.
REFERENCES
Arce, H., & Goodman, A. 1999, ApJ, 512, L135
Allain, S. 1998, A&A, 333, 629
Attridge, J. M., & Herbst, W. 1992, ApJ, 398, L61
Barrado y Navascu´es, D., Stauffer, J. R., & Patten B. M. 1999, ApJ, 522, 53
Bessell, M. S., Castelli, F., & Plez, B. 1998, A&A, 337, 321
Boesgaard, A. M., Budge, K., & Burck, E. E. 1988, ApJ, 325, 749
Boesgaard, A. M., Budge, K., & Ramsay, M. E. 1988, ApJ, 327, 389
Boesgaard, A. M., & Friel, E. D. 1990, ApJ, 351, 467
Cameron, A. C., & Campbell, C. G. 1993, A&A, 274, 309
Cameron, L. M. 1985, A&A, 147, 39
Chaboyer, B., Demarque, P., & Pinsonneault, M. H. 1995, ApJ, 441, 865
Crawford, D., & Perry, C. 1976, AJ, 81, 419
Dachs, J., & Kabus, H. 1989, A&AS, 78, 25
Dean, J. F., Warren, P. R., & Cousins, A. W. J. 1978, MNRAS, 183, 569
Debernardi, Y. & North, P. 2001, A&A, 374, 204
Eggen, O. J. 1972, ApJ, 173, 63
-- 21 --
Gonz´alez, J. F., & Lapasset, E. 2000, AJ, 119, 2296
Gray, D. F. 1992, The Observation and Analysis of Stellar Photospheres, 2nd edition, Cam-
bridge University Press, Cambridge
Harnden, F. R., et al. 2001, ApJ, 547, L141
Hartmann, L., Hewett, R., Stahler, S., & Mathieu, R. 1986, ApJ, 309, 275
Herbst, W., Bailer-Jones, C. A. L., & Mundt, R. 2001, ApJ, 554, L197
Herbst, W., Rhode, K. L., Hillenbrand, L. A., & Curran, G. 2000, AJ, 119, 261
Houdashelt, M. L., Bell, R. A., & Sweigart, A. V. 2000, AJ, 119, 1448
James, D. J., & Jeffries, R. D. 1997, MNRAS, 292, 252
Jeffries, R. D. 1997, MNRAS, 288, 585
Jeffries, R. D., James, D. J., & Thurston, M. R. 1998, MNRAS, 300, 550
Jeffries, R. D., Thurston, M. R., & Hambly, N. C. 2001, A&A, 375, 863
Jeffries, R. D., Thurston, M. R., & Pye, J. P. 1997, MNRAS, 287, 350
Jones, B. F., Fischer, D. A., & Stauffer, J. R. 1996, AJ, 112, 1562
Jones, B. F., Fischer, D., Shetrone, M., & Soderblom, D. R. 1997, AJ, 114, 352
Jones, B. F., & Prosser, C. F. 1996, AJ, 111, 1193
Kawaler, S. D. 1998, ApJ, 333, 236
Keppens, R., MacGregor, K. B., & Charbonneau, P. 1995, A&A, 294, 469
Kim, Y.-C., & Demarque, P. 1996, ApJ, 457, 340
Koester, D. & Reimers, D. 1996, A&A, 313, 810
Konigl, A. 1991, ApJ, 370, L39
Kraft, R. P. 1965, ApJ, 142, 681
Kraft, R. P. 1967, ApJ, 150, 551
Krishnamurthi, A., et al. 1998, ApJ, 493, 914
-- 22 --
Krishnamurthi, A., Pinsonneault, M. H., Barnes, S., & Sofia, S. 1997, ApJ, 475, 604
Kurucz, R. L. 1993, Kurucz CD-Rom 13, Atlas 9, SAO, Cambridge
Landolt, A. U. 1992, AJ, 104, 340
Lejeune, Th., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229
Lynga, G. 1987, Technical Report, Catalogue of Open Cluster Data., 5th ed., Lund Obser-
vatory
Lynga, G., & Wramdemark, S. 1984, A&A, 132, 58
MacGregor, K. B., & Brennan, M. 1991, ApJ, 376, 204
Meynet, G., Mermilliod, J.-C., & Maeder, A. 1993, A&AS, 98, 477
Micela, G., Sciortino, S., Jeffries, R. D., Thurston, M. R., & Favata, F. 2000, A&A, 357, 909
Milford, P. N., O'Mara, B. J., & Ross, J. E. 1994, A&A, 292, 276
Mills, D., Webb, J., & Clayton, M. 1997, Starlink User Note 152.4, Rutherford-Appleton
Laboratories
Narayanan, V. K., & Gould, A. 1999, ApJ, 523, 328
Nissen, P. E. 1988, A&A, 199, 146
Perryman, M. A. C., Brown, A. G. A., Lebreton, Y., Gomez, A., Turon, C., de Strobel, G.
C., Mermilliod, J.-C., Robichon, N., Kovalevsky, J., & Crifo, F. 1998, A&A, 331, 81
Pinsonneault, M. H., DePoy, D. L., & Coffee, M. 2001, ApJ, 556, 59
Pinsonneault, M. H., Stauffer, J., Soderblom, D. R., King, J. R., & Hanson, R. B. 1998,
ApJ, 504, 170
Pinsonneault, M. H., Terndrup, D. M., Stauffer, J. R., & Hanson, R. B. 2002, in preparation
Pinsonneault, M. H., Terndrup, D. M., & Yuan, Y. 2000, in Stellar Clusters and Associations:
Convection, Rotation, and Dynamics, ed. R. Pallavicini, G. Micela, & S. Sciortino,
ASP Conf. Ser. 198, 95
Queloz, D., Allain, S., Mermilliod, J.-C., Bouvier, J., & Mayor, M. 1998, A&A, 335, 183
Radick, R. R., Thompson, D. T., Lockwood, G. T., Duncan, D. K., & Baggett, W. E. 1987,
ApJ, 321, 459
-- 23 --
Rebull, L. M. 2001, AJ, 121, 1676
Robichon, N., Arenou, F., Mermilliod, J.-C., & Turon, C. 1999, A&A, 345, 471
Saxner, M., & Hammarback, G. 1985, A&A, 151, 372
Schatzman, E. 1962, Ann. d'Ap., 25, 18
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Sciortino, S., et al. 2001, A&A, 365, L259
Seiss, L., Dufour, E., & Forestini, M. 2000, A&A, 358, 593
Shu, F. H., Najita, J., Ruden, S. P., & Lizano, S. 1994, ApJ, 429, 797
Sills, A., Pinsonneault, M. H., & Terndrup, D. M. 2000, ApJ, 534, 335
Smith, K. C. 1992, PhD thesis, University of London
Soderblom, D. R., Jones, B. F., & Fischer, D. 2001, ApJ, 563, 334
Soderblom, D. R., Stauffer, J. R., Hudon, J. D., & Jones, B. F. 1993, ApJS, 83, 315
Stanek, K. Z. 1996, ApJ, 460, L37
Stassun, K. G., Mathieu, R. D., Mazeh, T., & Vrba, F. J. 1999, AJ, 117, 2941
Stauffer, J. R., et al. 1999, ApJ, 527, 219
Stauffer, J. R., Balachandran, S., Krishnamurthi, A., Pinsonneault, M. H., Terndrup, D. M.,
& Stern, R. A. 1997, ApJ, 475, 604
Stauffer, J. R., & Hartmann, L. 1987, ApJ, 318, 337
Stauffer, J. R., Hartmann, L. W., & Jones, B. F. 1989, ApJ, 346, 160
Stauffer, J. R., Hartmann, L. W., & Latham, D. W. 1987, ApJ, 320, L51
Stauffer, J. R., Jones, B. F., Backman, D., Pinsonneault, M., Hartmann, L. W., Barrado y
Navascu´es, D., & Terndrup, D. M. 2002, in preparation
Stauffer, J. R., Shultz, G., & Kirkpatrick, J. D. 1998, ApJ, 499, 199
Stover, R. J. 1998, in Instrumentation for Ground-Based Optical Astronomy, Present and
Future, ed. L. B. Robinson, (New York: Springer-Verlag), 443
-- 24 --
Stefanik, R. P., Latham, D. W., & Torres, G. 1999, in Precise Stellar Radial Velocities, IAU
Colloquium 170, eds. J. B. Hearnshaw & C. D. Scarfe, ASP Conf. Ser. 185, 345
Sung, H., Bessell, M. S., Lee, B.-W., & Lee, S.-G. 2002, AJ, 123, 290
Terndrup, D. M., Stauffer, J. R., Pinsonneault, M. H., Sills, A., Yuan, Y., Jones, B. F.,
Fischer, D., & Krishnamurthi, A. 2000, AJ, 119, 1303
Tinker, J., Pinsonneault, M. H., & Terndrup, D. M. 2002, ApJ, 564, 877
Twarog, B. A., Ashman, K. M., & Anthony-Twarog, B. J. 1997, AJ, 114, 2556
Udry, S., Mayor, M., & Queloz, D. 1999, in Precise Stellar Radial Velocities, IAU Colloquium
170, eds. J. B. Hearnshaw & C. D. Scarfe, ASP Conf. Ser. 185, 367
von Braun, K., & Mateo, M. 2001, AJ, 121, 1522
von Hippel, T., & Gilmore, G. 2000, AJ, 120, 1384
Weber, E. J., & Davis, L. 1967, ApJ, 148, 217
Weidemann, V. 2000, A&A, 363, 647
This preprint was prepared with the AAS LATEX macros v5.0.
-- 25 --
50
40
30
20
10
0
0
10
20
30
40
50
v sin i (JJT98)
i
)
k
r
o
w
s
h
t
(
i
n
s
v
i
Fig. 1. -- Comparison of new and previously published values of v sin i. The ordinate displays
v sin i values from this work, while the abscissa shows those measured by Jeffries et al. (1998).
The 1σ error bars are also shown. The discrepant point is for the star DK 206, discussed in
the text.
-- 26 --
6
8
10
12
14
0
V
−.5
0
.5
1
1.5
(B − V)0
Fig. 2. -- Color-magnitude diagram in V , B − V for NGC 2516. Filled points are for radial
velocity members from this survey, while the open points are for bright members in the
study of Gonz´alez & Lapasset (2000). Plusses denote non-members from this paper. The
photometry has been dereddened using E0(B − V ) = 0.12 and the extinction law derived
by Bessell, Castelli, & Plez (1998). The dashed line shows a 140 Myr isochrone, empirically
calibrated as described in the text, shifted to V0 − MV = 7.70, the Hipparcos distance to
NGC 2516 (Robichon et al. 1999). The solid line shows the solution from this paper, namely
(V0 − MV ) = 7.93 and [Fe/H] = −0.05. The arrow denotes the reddening vector.
-- 27 --
11
12
0
V
13
14
15
0
.5
1
1.5
(V − I)0
Fig. 3. -- Color-magnitude diagram in V , V − I for NGC 2516. Symbols and isochrones are
the same as in Fig. 2. Note the change in scale with respect to Fig. 2.
V
M
−
0
V
V
M
−
0
V
-- 28 --
V − I
B − V
10
12
V0
14
16
8.5
8
7.5
7
8.5
8
7.5
7
Fig. 4. -- Differences between the dereddened V magnitude and an empirically calibrated
isochrone for [Fe/H] = −0.05. The top panel shows the differences with respect to the
isochrone in V − I, while the lower panel shows these in B − V . The dashed line shows the
best fitting distance modulus V0 − MV = 7.93.
-- 29 --
)
s
/
m
k
(
i
i
n
s
v
100
10
100
10
100
10
100
10
a)
b)
c)
d)
.2
.4
.6
.8
1
1.2
1.4
(B − V)0
Fig. 5. -- Comparison of v sin i for (a) Pleiades, (b) NGC 2516, (c) M 34, (d) Hyades. Sources
for the rotation rates are described in the text. The solid line represents the mean Hyades
relation, while the dashed line characterizes the fast rotators in the Pleiades; these lines are
the same in all panels. Filled points denote measured v sin i values, while open points show
v sin i derived from photometric rotation periods. Upper limits to v sin i are displayed as
triangles.
-- 30 --
Fig. 6. -- Rotation rate as a function of effective temperature for stars in young open clusters
with ages between 100 Myr and 600 Myr. The assumed ages of the clusters are shown in
each figure. The solid points are detections and the open triangles are upper limits. The
solid lines show the five different values of the disk-locking lifetime that were used: 0, 0.3,
1, 3, and 10 Myr (from top to bottom).
)
l
o
b
L
/
X
L
(
g
o
l
−2
−3
−4
−5
1
-- 31 --
10
100
τ
conv / Prot
1000
Fig. 7. -- X-ray luminosity against inverse Rossby number. Solid points show stars in
NGC 2516, while open points are Pleiades stars with measured periods. Triangles show
upper limits. See text for details.
-- 32 --
Table 1. NGC 2516 data
Star
Source
v(r)
σ(v)
v sin i
σ
V
B − V
V − I
JTH phot?
CTIO-1
CTIO-2
CTIO-3
CTIO-5
CTIO-6
CTIO-7
DK078
DK081
DK206
DK215
DK221
DK230
DK232
DK233
DK234
DK245
DK307
DK308
DK311
DK313
DK320
DK323
DK325
DK403
DK409
DK415
DK428
DK463
DK513
DK514
DK526
2
2
2
2
2
2
1
2
1,2
1
2
1
3
1
1
1
3
1,2
1
1,2
1,2
1
1,2
1
2
2
1
2
1
2
1
24.2
23.3
24.7
24.8
22.1
23.7
27.4
26.4
23.2
25.9
23.4
27.1
24.2
22.8
20.0
24.8
21.2
23.9
24.0
24.8
24.8
23.7
23.3
30.4
22.5
24.8
22.2
25.5
23.5
24.1
24.9
1.0
0.6
0.6
0.8
1.6
0.7
0.9
5.0
1.0
0.8
0.8
0.5
0.6
0.6
0.7
0.4
2.1
1.3
3.1
0.9
0.8
0.8
0.5
1.2
0.7
1.5
0.5
0.6
1.2
0.6
1.1
≤ 8
≤ 6
≤ 6
≤ 7
11.7
10.5
24.2
· · ·
18.3
21.6
≤ 7
12.8
10.0
12.6
9.2
9.5
73.8
10.5
37.7
15.4
13.7
22.6
11.6
26.3
≤ 7
15.3
16.4
7.5
24.4
≤ 7
25.4
· · ·
· · ·
· · ·
· · ·
3.0
1.1
2.2
· · ·
4.6
1.2
· · ·
1.0
7.0
1.0
1.2
1.1
1.6
1.6
3.7
1.3
1.1
2.1
0.8
1.7
· · ·
2.4
0.7
1.0
2.1
· · ·
1.9
14.701
14.184
14.287
14.804
13.590
13.735
12.068
10.725
12.696
12.682
13.293
12.792
11.370
12.240
13.661
13.541
11.753
13.047
12.437
13.289
12.878
12.503
13.020
12.905
13.730
12.954
13.312
13.399
12.647
13.839
12.759
1.014
0.903
0.938
1.051
0.779
0.804
0.603
0.328
0.637
0.636
0.814
0.597
0.386
0.674
0.874
0.768
0.499
0.684
0.602
0.732
0.675
0.592
0.671
0.719
0.809
0.671
0.751
0.742
0.655
0.831
0.649
1.094
1.007
1.007
1.105
0.889
0.896
0.741
0.445
0.743
0.750
0.875
0.704
0.474
0.769
0.987
0.853
0.625
0.794
0.718
0.843
0.818
0.701
0.762
0.876
0.907
0.800
0.924
0.873
0.802
0.922
0.778
6049 Y
6263 Y
6676 Y
8654 Y
9361 Y
10753 Y
7864 Y
8058 Y
7967 Y
7104 Y
6880 Y
6700 N
6446 Y
6029 Y
5887 Y
4914 Y
7189 Y
6452 Y
5833 Y
5862 Y
3208 Y
3621 Y
4574 Y
7585 Y
6852 Y
5901 Y
4105 Y
6570 Y
7743 Y
7312 Y
6976 Y
-- 33 --
Table 1 -- Continued
Star
Source
v(r)
σ(v)
v sin i
σ
V
B − V
V − I
JTH phot?
DK609
DK612
DK664
DK668
DK722
DK742
DK803
DK820
DK875
DK902
DK904
DK907
DK908
DK914
DK933
E010
E020
E027
E063
E095
GSC89111169
GSC89111238
GSC89111475
GSC89113284
JTH 758
JTH 14223
1
3
1
1
1
3
2
1
3
2
1
3
2
2
1
2
1
2
1
1
1,2
1
1
1
3
3
24.4
25.1
25.9
19.6
24.5
22.5
23.2
22.6
25.6
20.8
27.6
25.1
24.7
24.5
23.1
23.2
25.1
27.1
24.9
25.0
22.1
24.5
24.6
22.6
23.7
23.0
0.4
0.5
1.8
1.8
0.6
0.7
1.1
0.4
1.3
5.0
1.0
1.0
0.7
0.6
0.5
2.7
0.4
1.8
0.4
0.4
2.7
1.1
1.2
0.5
0.7
0.7
10.9
65.5
47.9
42.8
8.6
53.1
16.4
9.7
108.8
· · ·
22.6
24.4
10.1
9.1
14.1
32.4
9.7
19.4
8.7
7.6
38.5
27.2
37.1
11.8
24.4
40.6
1.0
4.7
1.6
1.2
0.9
1.0
2.2
1.2
8.4
· · ·
1.0
0.5
1.1
1.0
0.8
3.3
1.3
2.4
1.1
1.2
2.9
1.3
1.4
1.2
2.0
1.4
12.142
11.534
13.840
14.120
13.756
11.704
13.405
14.279
12.000
11.465
12.841
12.336
12.871
13.641
12.607
14.335
14.322
14.429
14.509
14.240
14.516
13.984
14.576
14.867
12.632
12.121
Nonmembers
0.659
0.444
0.818
1.065
0.801
0.480
0.736
0.934
0.526
0.533
0.691
0.583
0.660
0.780
0.762
0.903
0.935
0.928
0.976
0.930
0.943
0.861
1.015
1.034
0.640
0.537
0.758
0.542
0.963
1.141
0.896
0.795
0.838
1.085
0.647
0.665
0.836
0.709
0.760
0.851
0.864
1.048
1.008
1.052
1.061
1.000
1.118
1.016
1.080
1.171
0.794
0.672
9852 Y
10301 Y
9465 Y
8645 Y
11716 Y
14681 N
9140 Y
12118 Y
12302 Y
8458 Y
9054 Y
8529 Y
8536 Y
9446 Y
9486 Y
9676 Y
7619 Y
7678 Y
10871 Y
9286 Y
8634 Y
4598 Y
5957 Y
10040 Y
· · · Y
· · · Y
DK306
DK365
1
1
82.0
91.3
6.9
53.1
0.8 ≤ 7.0
2.5
· · ·
13.747
13.968
0.726
1.022
0.856
1.108
7147 N
6705 Y
-- 34 --
Table 1 -- Continued
Star
Source
v(r)
σ(v)
v sin i
σ
V
B − V
V − I
JTH phot?
DK417
DK503
DK508
DK573
DK574
DK625
DK669
DK813
DK919
DK967
E013
E111
1
1
2
1
1
1
1
1
1
2
1
1
9.0
65.6
43.6
0.8
77.0
337.0
-- 0.4
38.3
13.8
56.0
31.2
5.1
17.9
1.1
32.4
1.9
126.3
2.6
≤ 7
0.7
≤ 7
0.6
≤ 7
0.4
≤ 7
0.5
8.8
0.5
7.3
0.3
10.0
0.5
0.5
9.9
0.5 ≤ 7.0
1.4
3.4
5.2
· · ·
· · ·
· · ·
· · ·
0.5
1.3
7.0
1.3
· · ·
12.053
12.706
11.250
14.380
14.009
13.892
13.933
12.029
14.215
12.164
14.762
14.037
0.515
0.620
0.376
0.874
1.043
0.822
0.966
0.654
0.950
0.546
0.996
0.817
0.619
0.768
0.456
1.048
1.112
0.999
1.046
0.980
0.982
0.745
1.139
0.979
6096 Y
7650 Y
6949 Y
6307 N
6724 Y
11031 Y
8689 Y
10390 Y
8262 Y
12923 Y
8967 Y
6649 Y
Note. -- Sources for the data: (1) Observations at CTIO (this study); (2) Jeffries et al.
(1998); (3) Observations at AAT (this study).
-- 35 --
Table 2. Cluster radial velocity
Sample
r.v. (km s−1) N
All stars (this study)
Jeffries et al. (1998)
Gonz´alez & Lapasset (2000)
24.2 ± 0.2
23.8 ± 0.3
22.0 ± 0.2
57
24
22
Table 3. Error budget for main-sequence fit
Source of error
value ∆(V0 − MV ) ∆([Fe/H])
Fit to data points
Error in E(B − V )
±0.02
Error in E(V − I)/E(B − V ) ±0.02
Total error
· · ·
· · ·
±0.015
∓0.020
∓0.065
∓0.14a
±0.02
±0.10
±0.06
±0.14
aIncluding the effect of the abundance error, which is ∆(V0 −
MV )/∆[Fe/H] = +1.3.
|
0805.1709 | 1 | 0805 | 2008-05-12T18:42:15 | Galaxy Evolution in Hickson Compact Groups: The Role of Ram Pressure Stripping and Strangulation | [
"astro-ph"
] | Galaxies in compact groups tend to be deficient in neutral hydrogen compared to isolated galaxies of similar optical properties. In order to investigate the role played by a hot intragroup medium (IGM) for the removal and destruction of HI in these systems, we have performed a Chandra and XMM-Newton study of eight of the most HI deficient Hickson compact groups. Diffuse X-ray emission associated with an IGM is detected in four of the groups, suggesting that galaxy-IGM interactions are not the dominant mechanism driving cold gas out of the group members. No clear evidence is seen for any of the members being currently stripped of any hot gas, nor for galaxies to show enhanced nuclear X-ray activity in the X-ray bright or most HI deficient groups. Combining the inferred IGM distributions with analytical models of representative disc galaxies orbiting within each group, we estimate the HI mass loss due to ram pressure and viscous stripping. While these processes are generally insufficient to explain observed HI deficiencies, they could still be important for HI removal in the X-ray bright groups, potentially removing more than half of the ISM in the X-ray bright HCG 97. Ram pressure may also have facilitated strangulation through the removal of galactic coronal gas. In X-ray undetected groups, tidal interactions could be playing a prominent role, but it remains an open question whether they can fully account for the observed HI deficiencies. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 21 (2008)
Printed 1 June 2021
(MN LATEX style file v2.2)
Galaxy evolution in Hickson compact groups: The role of ram
pressure stripping and strangulation
Jesper Rasmussen,1⋆† Trevor J. Ponman,2 Lourdes Verdes-Montenegro,3 Min S. Yun4
and Sanchayeeta Borthakur4
1Observatories of the Carnegie Institution, 813 Santa Barbara Street, Pasadena, CA 91101, USA
2School of Physics and Astronomy, University of Birmingham, Edgbaston, Birmingham B15 2TT
3Instituto de Astrof´ısica de Andaluc´ıa, CSIC, Apdo. Correos 3004, E-18080 Granada, Spain
4Astronomy Department, University of Massachusetts, Amherst, MA 01003, USA
ABSTRACT
Galaxies in compact groups tend to be deficient in neutral hydrogen compared to isolated
galaxies of similar optical properties. In order to investigate the role played by a hot intragroup
medium (IGM) for the removal and destruction of HI in these systems, we have performed a
Chandra and XMM-Newton study of eight of the most HI deficient Hickson compact groups.
Diffuse X-ray emission associated with an IGM is detected in four of the groups, suggesting
that galaxy -- IGM interactions are not the dominant mechanism driving cold gas out of the
group members. No clear evidence is seen for any of the members being currently stripped of
any hot gas, nor for galaxies to show enhanced nuclear X-ray activity in the X-ray bright or
most HI deficient groups. Combining the inferred IGM distributions with analytical models
of representative disc galaxies orbiting within each group, we estimate the HI mass loss due
to ram pressure and viscous stripping. While these processes are generally insufficient to
explain observed HI deficiencies, they could still be important for HI removal in the X-ray
bright groups, potentially removing more than half of the ISM in the X-ray bright HCG 97.
Ram pressure may also have facilitated strangulation through the removal of galactic coronal
gas. In X-ray undetected groups, tidal interactions could be playing a prominent role, but it
remains an open question whether they can fully account for the observed HI deficiencies.
Key words: galaxies: evolution -- galaxies: interactions -- galaxies: ISM -- X-rays: galaxies
-- X-rays: galaxies: clusters
1 INTRODUCTION
The origin of the galaxy morphology-density relation is still one
of the most important unsolved problems in astrophysics. Not only
are spiral galaxies less common within dense cluster environments,
but those which are present tend to be deficient in HI, and this de-
ficiency itself correlates with projected local galaxy density (e.g.,
Giovanelli & Haynes 1985). The mechanisms responsible for the
changes in the morphology and gas content of galaxies are unclear,
with gas stripping, tidal shocks, and galaxy interactions and merg-
ers all contenders.
Traditionally, clusters of galaxies have represented the envi-
ronment of choice for attempts to unravel the nature of the rele-
vant processes. Recent results, however, strongly suggest that the
origin of the environmental modification of galaxies which under-
pin the morphology-density relation lies, not in the cores of galaxy
clusters, but in smaller groups and cluster outskirts. For exam-
⋆ E-mail: [email protected]
† Chandra Fellow
ple, spectroscopic studies of the effects of the cluster environment
on galaxies (e.g., Lewis et al. 2002) show that the suppression of
star formation takes place in cluster outskirts rather than in the
core, and is modulated by local galaxy density. Moreover, X-ray
bright groups show a morphology-density relation stronger than
that of clusters (Helsdon & Ponman 2003), a result adding to the
accumulating evidence that cluster galaxies have often been 'pre-
processed' in groups prior to their assembly into larger systems
(see, e.g., Cortese et al. 2006). In addition, it is becoming clear that
processes once thought to be exclusive to the cluster environment,
such as ram pressure stripping (Rasmussen, Ponman & Mulchaey
2006) and strangulation (Kawata & Mulchaey 2008), may play a
role also in much smaller systems. In order to elucidate the ori-
gin of the morphology-density relation, it is therefore necessary to
study the processes acting on galaxies within groups.
The compact groups in the catalogue of Hickson (1982) of-
fer particularly interesting opportunities in this regard. Many of
these groups are spiral-rich, but their galaxy population is, on av-
erage, deficient in HI by a factor of ∼ 2 compared to loose groups
(Williams & Rood 1987). Furthermore, some of these groups have
2
J. Rasmussen et al.
exceptionally compact galaxy configurations, and so may repre-
sent pre-virialisation systems close to maximum collapse, in which
galaxies are suffering strong environmental modification but have
yet to be converted into early-types. While recent work has iden-
tified tidal interactions and mergers as playing an important role
in the morphological transformation of spirals in some compact
groups (Coziol & Plauchu-Frayn 2007), it is still unclear what is
causing the observed HI deficiencies and to what extent this is re-
lated to the processes modifying the stellar component of the group
members.
Verdes-Montenegro et al. (2001) presented a detailed study of
the HI content of a sample of 72 Hickson compact groups (HCGs),
with integrated HI masses from single dish measurements, and de-
tailed Very Large Array (VLA) mapping of a subsample. Defining
HI deficiency ∆HI as
∆HI ≡ logMHI,pred − logMHI,obs,
(1)
where MHI,obs is the observed HI mass of the group galaxies and
MHI,pred is that predicted for isolated galaxies of similar morphol-
ogy and optical luminosity (Haynes & Giovanelli 1984), this work
confirmed the deficiency in HCGs, and allowed a search for corre-
lations between deficiency and other group properties. One of the
strongest relationships found was that with detectable intergalactic
X-ray emission; almost half of the groups with significant HI de-
ficiency showed diffuse intragroup X-ray emission in the ROSAT
survey of Ponman et al. (1996, hereafter P96). More recently, sim-
ilar results have been reported also for groups outside Hickson's
catalogue (Sengupta & Balasubramanyam 2006).
The increased prevalence of X-ray detected systems among
HI deficient (compact) groups may suggest a picture whereby
HI is stripped from spiral galaxies within virialising groups, and
then destroyed due to heating by a surrounding hot intragroup
medium (IGM). However, for most of the compact groups in the
Verdes-Montenegro et al. (2001) sample, ROSAT data were either
not available or of insufficient quality to permit any detailed study
of the relationship between the hot and cold gas. With only very
shallow ROSAT All-Sky Survey data at their disposal for a number
of these systems, Verdes-Montenegro et al. (2001) could not estab-
lish the amount and properties of any hot IGM in these groups.
In order to explore the processes destroying HI in these sys-
tems, we have therefore embarked on a programme to obtain high
quality X-ray and radio data for the most HI deficient compact
groups, adding in existing archival X-ray data wherever relevant.
Since ram pressure stripping can only operate where a significant
IGM is present, while tidal stripping of gas requires only that galax-
ies be interacting, a key discriminator for the mechanism of HI re-
moval from galaxies is whether or not there is a correlation between
HI deficiency and the properties of a hot IGM, and in particular
whether a hot IGM is present in the highly HI deficient systems.
Clarifying these issues represents the goal of the present study, in
an attempt to shed light on the role played by galaxy -- IGM inter-
actions in destroying HI and establishing the morphology-density
relation in these compact groups. The focus of this paper thus
rests mainly on the X-ray properties of any hot intergalactic gas
within the groups, while a forthcoming companion paper (Verdes-
Montenegro et al., in prep.) will discuss in more detail the HI and
radio continuum emission in the groups, the detailed relationship
between X-ray and HI morphology, and the multi-wavelength prop-
erties of the individual group galaxies.
In Sections 2 and 3 we outline the sample selection and X-
ray data analysis, respectively. Section 4 presents the results for
the X-ray properties of the hot gas and galaxies within each group,
and compares the derived IGM properties to the observed HI defi-
ciencies. In Section 5 we construct an analytical model of a repre-
sentative late-type galaxy orbiting within the derived gravitational
potential of each X-ray bright group, allowing us to evaluate the
importance of ram pressure and viscous stripping for HI removal.
The results are discussed in Section 6 and summarized along with
the main conclusions in Section 7.
A Hubble constant of 73 km s−1 Mpc−1 is assumed through-
out. Unless otherwise stated, all errors are quoted at the 68 per cent
confidence level.
2 GROUP SAMPLE AND OBSERVATIONS
Our broad aim is to establish the relationship between the hot and
cold gas in HI deficient groups, and explore the processes of gas
removal operating within them. Our sample is therefore based on
that studied by Verdes-Montenegro et al. (2001), from which we
selected all HCGs which are highly deficient in HI (∆HI > 0.5)
based on Very Large Array (VLA) HI data, contain four or more
group galaxies, and lie at a distance D < 100 Mpc. This yielded an
initial sample of 11 groups, which will be discussed in its entirety
in Verdes-Montenegro et al. (in prep.). VLA HI mapping has been
completed for all 11 systems, along with follow-up Green Bank
Telescope (GBT) observations for the groups discussed in this pa-
per. Compared to the VLA, the single-dish GBT is more sensitive
to extended, smoothly distributed emission that may otherwise be
filtered out by the VLA interferometer. Including the GBT data thus
provides a more complete census of the HI content in the groups.
For HCG 48, included in the initial sample, the additional gas de-
tected by GBT indicates that this group is not HI deficient after
all. Consequently, this group was omitted from the sample for the
purpose of the present study.
Of the remaining systems, the eight included in this paper are
those for which Chandra or XMM-Newton X-ray data are currently
available. Four of these eight groups are exceptionally compact,
with major galaxies (as catalogued by Hickson 1982) contained
within a circle of radius ∼ 2 arcmin, and thus requiring Chan-
dra data to resolve their X-ray structure in the crucial region. We
note that these groups are not generally mature, X-ray bright sys-
tems dominated by early-type galaxies. Such groups may contain
little HI, but the HI content of their galaxies as predicted from the
Haynes & Giovanelli (1984) results mentioned above is also small
(zero for ellipticals and modest for lenticulars). Such groups are
therefore not usually HI deficient according to our definition in
equation (1). Our chosen systems have HI content far below that
expected for their galaxy contents, and are therefore those in which
the processes which destroy HI should be in active, or very recent,
operation.
The observation log for the X-ray data considered in this paper
is presented in Table 1, detailing the pointing coordinates for each
observation (for archival observations not necessarily identical to
the optical group centre), group distance D, the observing instru-
ment, date, and mode, along with cleaned exposure times texp for
each camera, and the Galactic absorbing column density NH from
Dickey & Lockman (1990) as adopted in the X-ray spectral analy-
sis.
Ram pressure stripping in Hickson groups
3
Table 1. Log of available X-ray observations. Group luminosity distances D for the adopted value of H0 were taken from
the NASA/IPAC Extragalactic Database (NED). Column 7 specifies the frame mode (full frame/extended full frame) and
optical blocking filter for XMM, and ACIS CCD aimpoint and telemetry mode (Faint/Very Faint) for Chandra.
Group
RA
(J2000)
Dec
(J2000)
D
(Mpc)
HCG 7
. . .
. . .
HCG 15
. . .
. . .
HCG 30
HCG 37
HCG 40
. . .
HCG 44
. . .
. . .
. . .
HCG 97
HCG 100
. . .
00 39 13.5 +00 51 49.3
. . .
. . .
. . .
. . .
02 07 39.0 +02 08 18.0
. . .
. . .
. . .
. . .
04 36 28.3 −02 50 02.9
09 13 35.3 +30 01 25.5
09 38 56.7 −04 50 55.9
. . .
. . .
10 18 02.5 +21 48 50.7
10 17 38.0 +21 41 17.0
. . .
. . .
. . .
. . .
23 47 25.4 −02 19 45.5
00 01 25.9 +13 06 30.4
. . .
. . .
54
. . .
. . .
92
. . .
. . .
63
97
98
. . .
23
. . .
. . .
. . .
86
69
. . .
Chandra/
XMM
XMM pn
XMM m1
XMM m2
XMM pn
XMM m1
XMM m2
Chandra
Chandra
Chandra
. . .
Chandra
XMM pn
XMM m1
XMM m2
Chandra
Chandra
. . .
Obs. date
Obs. mode
(yyyy-mm-dd)
2004-12-26
. . .
. . .
2002-01-10
. . .
. . .
2006-02-07
2005-01-13
2005-01-29
2005-01-29
2002-03-14
2001-05-07
. . .
. . .
2005-01-14
2006-06-12
2007-01-24
FF Thin
FF Thin
FF Thin
EFF Thin
FF Thin
FF Thin
ACIS-S VF
ACIS-S VF
ACIS-S VF
ACIS-S VF
ACIS-S F
EFF Thick
FF Thin
FF Thin
ACIS-S VF
ACIS-I VF
ACIS-I VF
texp
(ks)
25.9
35.1
35.9
23.3
31.0
31.1
29.0
17.5
31.8
14.6
19.4
8.1
13.4
13.1
36.2
26.1
16.1
NH
(1020 cm−2)
2.24
. . .
. . .
3.20
. . .
. . .
5.08
2.30
3.64
. . .
2.16
. . .
. . .
. . .
3.65
4.40
. . .
3 DATA REDUCTION AND ANALYSIS
3.1 XMM-Newton data
The XMM data were analysed using XMMSAS v6.5.0, and cali-
brated event lists were generated with the 'emchain' and 'epchain'
tasks. Event files were filtered using standard quality flags, while
retaining only patterns 6 4 for pn and 6 12 for MOS. Screening
for background flares was first performed in the 10 -- 15 keV band
for MOS and 12 -- 14 keV for pn. Following an initial removal of
obvious large flares, a 3σ clipping of the resulting lightcurve was
applied. Point sources were then identified by combining the re-
sults of a sliding-cell search ('eboxdetect') and a maximum likeli-
hood point spread function fitting ('emldetect'), both performed in
five separate energy bands spanning the total range 0.3 -- 12 keV. In
order to filter out any remaining soft protons in the data, a second
lightcurve (3σ) cleaning was then done in the 0.4 -- 10 keV band,
within a 9 -- 12 arcmin annulus which excluded the detected point
sources. Closed-filter data from the calibration database and blank-
sky background data (Read & Ponman 2003) for the appropriate
observing mode were filtered similarly to source data, and screened
so as to contain only periods with count rates within 1σ from the
mean of the source data. All point sources were excised out to at
least 25 arcsec in spectral analysis.
To aid the search for diffuse X-ray emission within the groups,
smoothed exposure-corrected images were produced, with back-
ground maps generated from blank-sky data. We allowed for a
differing contribution from the non-vignetted particle background
component in source- and blank-sky data by adopting the following
approach. First an EPIC mosaic image was smoothed adaptively
(3σ -- 5σ significance range), and the particle background was sub-
tracted. The latter was estimated from closed-filter data which were
scaled to match source data count rates in the image energy band
in regions outside the field of view, and then smoothed at the same
spatial scales as the source data. The resulting photon image in-
cludes the X-ray background at the source position. To remove this
component, a particle-subtracted blank-sky image was produced in
a similar way, and scaled to match 0.3 -- 2 keV source count rates
in a point-source -- excised 10 -- 12 arcmin annulus assumed to be
free of IGM emission (this assumption is clearly justified in all
cases, as will be shown). This image was then subtracted from the
corresponding source image, and the result was finally exposure-
corrected.
3.2 Chandra data
For all Chandra data sets, calibrated event lists were regenerated
using CIAO v3.3. For Very Faint mode observations, the standard
additional background screening was carried out. Bad pixels were
screened out using the bad pixel map provided by the pipeline,
and remaining events were grade filtered, excluding ASCA grades
1, 5, and 7. Periods of high background were filtered using 3σ
clipping of full -- chip lightcurves, binned in time bins of length
259.8-s and extracted in off-source regions in the 2.5 -- 7 keV band
for back-illuminated chips and 0.3 -- 12 keV for front-illuminated
chips. Blank-sky background data from the calibration database
were screened and filtered as for source data, and reprojected to
match the aspect solution of the latter. Point source searches were
carried out with the CIAO task 'wavdetect' using a range of scales
and detection thresholds, and results were combined. Source ex-
tents were quantified using the 4σ detection ellipses from 'wavde-
tect', and these regions were masked out in all spectral analysis.
In order to produce smoothed images as for the XMM data,
background maps were generated using blank-sky data, and scaled
to match source count rates for each CCD. This scaling employed
either the full-chip 10 -- 12 keV count rates, with point sources ex-
cluded, or, where possible, count rates in the image energy band
within source-free regions on the relevant CCD. The background
maps were then smoothed to the same spatial scales as the source
data and subtracted from the latter. The resulting images were
finally exposure corrected using similarly smoothed, spectrally
weighted exposure maps, with weights derived from spectral fits to
the integrated diffuse emission (where possible, otherwise the ex-
posure maps were weighted by a T = 1 keV, Z = 0.3 Z⊙ thermal
plasma model).
4
J. Rasmussen et al.
3.3 Spatial and spectral analysis
While the smoothed X-ray images described above are useful in
terms of establishing the presence and rough morphology of any
intragroup medium, we emphasize that they were used for illus-
trative purposes only and not for quantitative analyses. Where
IGM emission was not immediately obvious from these images,
we performed an additional source detection procedure based on
Voronoi tessellation and percolation ('vtpdetect' in CIAO), which
can be useful for detecting extended, low -- surface brightness emis-
sion missed by our standard detection algorithms. To reduce the
fraction of spurious detections, a minimum of 50 net counts were
required for a source to be considered real.
As a second step in the search for group-scale diffuse emis-
sion, we also extracted exposure-corrected 0.3 -- 2 keV surface
brightness profiles from the unsmoothed data, with all detected
point-like and extended galactic sources masked out. The profiles
were extracted from the optical group centre defined by the prin-
cipal members in the Hickson (1982) catalogue. For XMM data,
we used an EPIC mosaic image for this purpose, with the particle
background removed using the method described above. The es-
timated particle level shows a typical standard error of the mean
of ≈ 5 per cent, which should be representative of the uncertainty
associated with particle subtraction if the ratio of particle events
inside and outside the field of view is similar to that in the closed-
filter data. We used the blank-sky background data to confirm this
assumption (since these have very little source contamination), but
have conservatively added a 10 per cent error in quadrature to our
XMM surface brightness errors, to allow for any residual systematic
uncertainties associated with the particle subtraction.
For the spectral analysis of any extended emission, X-ray
spectra were accumulated in energy bins of at least 20 net counts,
and fitted in XSPEC v11.3 assuming an APEC thermal plasma
model with the solar abundance table of Grevesse & Sauval (1998).
XMM background spectra were extracted by means of the common
'double-subtraction' technique (Arnaud et al. 2002), using blank-
sky background data for the on-chip background, and a large-radius
(10 -- 12 arcmin) annulus for determining the local soft X-ray back-
ground. Owing to the smaller field of view, a similar approach was
not generally possible or desirable for the Chandra observations
where source emission may completely fill the CCD under inves-
tigation. The extraction of Chandra background data products are
therefore described individually for each group in the next Section.
Surface brightness profiles of the X-ray detected groups were
extracted from the peak of the diffuse X-ray emission when clearly
identifiable (in HCG 37 and 97) and from the centroid otherwise
(HCG 15 and 40). The profiles were fitted with standard β -- models
for conversion into IGM density profiles under the assumption of
isothermality. Since X-ray emissivity is very nearly independent of
temperature for a T ∼ 0.5 -- 1 keV plasma of the relevant metal-
licities (see Sutherland & Dopita 1993), this approach is entirely
adequate for our purposes, where the uncertainties of our analysis
will ultimately be dominated by those related to the modelling of
the impact of the hot gas on the galaxies. The density profiles were
normalized using the spectral normalization A from XSPEC,
A =
10−14
4πD2(1 + z)2 ZV
nenH dV cm−5,
(2)
where D is the assumed group distance, and ne and nH are the
number densities of electrons and hydrogen atoms, respectively.
The integral represents the fitted emission integral Ie over the cov-
ered volume V , assumed to be spherically symmetric. Total gas
masses within the region of interest were derived by simple vol-
ume integration of ne(r)µemp, where mp is the proton mass and
we have assumed ne/nH = 1.165 and µe = 1.15, appropriate for
a fully ionized Z = 0.3 Z⊙ plasma at the relevant temperatures
(Sutherland & Dopita 1993).
Since we have no knowledge of the density distribution of
any hot gas in the X-ray undetected groups, we have only derived
constraints on their mean electron density hnei within the region
considered, effectively assuming a uniform IGM distribution. The
advantage of this approach is that it provides very conservative up-
per limits to the IGM masses (and mean ram pressures) within the
relevant region. The derived limits to hnei and MIGM for these
groups were obtained from the IGM count rate limits and thus de-
pend on the depth of the X-ray data. These count rate limits were
established on the basis of the exposure-corrected (and, in the case
of XMM, particle-subtracted) images, by comparing the emission
level within a region centred on the optical group centre with that in
a surrounding annulus. The physical extent of the region of interest
was thus constrained by the need to evaluate the background locally
from our data, and varies from 50 -- 150 kpc among the groups, as
detailed in the discussion of individual groups below. The derived
constraints on IGM count rates were translated into constraints on
Ie, assuming the cooling function Λ(T, Z) of Sutherland & Dopita
(1993) and an IGM temperature taken from the σV -- TX relation of
Osmond & Ponman (2004),
log σV = (1.15 ± 0.26) log TX + 2.60 ± 0.03,
(3)
with galaxy velocity dispersions σV in km s−1 and TX in keV. Er-
rors on T were derived from the dispersion of this relation, with σV
taken from P96 for HCG 7, 30, 37, 44, and 100, from Mahdavi et al.
(2005) for HCG 97, and from Osmond & Ponman (2004) for the re-
mainder. The resulting temperature range was then used to estimate
3σ upper limits on hnei ∼ (Ie/V )1/2 inside the assumed spheri-
cal volume V for any subsolar metallicity Z. The assumption of a
uniform IGM in these groups implies that we can identify the up-
per limit to the central IGM density n0 with hnei and constrain the
IGM masses by simply multiplying hnei and V .
In order to briefly investigate the level of nuclear X-ray activ-
ity among individual group members, 0.3 -- 2 keV count rates of all
galactic central point sources were also extracted, adopting extrac-
tion regions of 2 and 15 arcsec radius for Chandra and XMM data,
respectively. In the majority of cases, photon statistics were insuf-
ficient to allow robust spectral fitting for individual sources. For
consistency, all point source count rates were therefore converted
to luminosities assuming a power-law spectrum of photon index
Γ = 1.7, absorbed by the Galactic value of NH. The associated
uncertainties were derived from the Poisson errors on the photon
count rates.
4 RESULTS
In this section we discuss the results obtained for the IGM in each
group and for the X-ray emission associated with individual group
galaxies. Figure 1 shows contours of the smoothed, background-
subtracted X-ray emission of each group overlayed on Digitized
Sky Survey (DSS) images. Diffuse X-ray emission associated with
an intragroup medium is detected in four of the eight groups, as de-
scribed for each group individually below. For the remaining four,
we do not detect any extended group emission, neither inside a
given physical radius from the optical group centre when compared
to the emission level in surrounding regions, nor on the basis of the
Ram pressure stripping in Hickson groups
5
Figure 1. Contours of adaptively smoothed 0.3 -- 2 keV emission overlayed on DSS images for all groups, with the principal group members labelled following
the notation of Hickson (1982). Where relevant, dashed squares outline the coverage of the Chandra/ACIS CCD's. For HCG 44, dark (black) contours outline
the Chandra emission, and lighter (green) those of an overlapping XMM pointing.
Voronoi source detection procedure. To corroborate these results,
Figures 2 and 3 show the derived surface brightness profiles for the
groups with and without detectable intragroup emission, respec-
tively.
Table 2 summarizes the observed HI and X-ray properties of
the groups, along with the adopted velocity dispersions from optical
spectroscopy. HI deficiencies in the Table are from our GBT mea-
surements (Borthakur et al., in prep.), except for HCG 44, for which
we have adopted the older VLA value (Verdes-Montenegro et al.
2001) due to the GBT beam size only covering the central region
of this relatively nearby system. The listed HI deficiencies are based
on the integrated HI mass within the circular region covered by the
radio data (rHI in the Table). In many groups, a significant fraction
of the detected HI is located outside the optical extent of individ-
ual galaxies (i.e. is intergalactic) and cannot be clearly assigned to
any individual group member. The derived values of ∆HI should
therefore generally be viewed as an average for the galaxies within
the GBT or VLA field. As the fractional 1-σ uncertainty on the
measured HI masses from our GBT data is less than 1 per cent for
all groups, uncertainties on the listed deficiencies are dominated by
those related to the predicted logarithmic HI mass, which we have
taken to be 0.2, adopting the standard estimate of error provided by
Haynes & Giovanelli (1984).
For reference, 0.3 -- 2 keV X-ray luminosities are also listed in
Table 2, corrected for Galactic absorption, and derived within the
region employed for the spectral analysis unless otherwise speci-
fied in the subsection for the relevant group. The listed central hot
IGM densities (or upper limits to the mean densities for the X-
ray undetected groups) were computed as outlined in Section 3.3.
IGM masses in the Table were derived within the same region as
the HI deficiencies (i.e. within rHI given in the Table), to enable a
direct comparison between the two. Note, as indicated above, that
6
J. Rasmussen et al.
Table 2. Summary of derived group properties. Except where indicated, HI deficiencies ∆HI are from GBT
data (Borthakur et al., in prep.), obtained inside a radius rHI. Column 7 lists the derived constraints on central
hot IGM density, with upper limits for the X-ray undetected groups given at 3σ significance, and Column 8
the corresponding hot IGM mass within rHI.
Group
∆HI
HCG 7
HCG 15
HCG 30
HCG 37
HCG 40
HCG 44
HCG 97
HCG 100
0.60
0.46
1.37
0.33
0.60
0.69†
0.35
0.27
rHI
(kpc)
σV
(km s−1)
TX
(keV)
LX
n0
(1040 erg s−1)
(10−3 cm−3)
68
113
79
119
121
101
106
86
95a
404b
72a
446a
157b
145a
383c
100a
0.3 ± 0.1∗
0.83+0.09
−0.06
0.2 ± 0.1∗
0.86+0.13
−0.09
0.59+0.11
−0.12
0.4 ± 0.1∗
0.97+0.14
−0.12
0.3 ± 0.1∗
< 0.5
32 ± 2
< 1.3
16 ± 2
3.1 ± 0.5
< 3.6
120+20
−24
< 3.4
< 0.06
4.4 ± 0.5
< 1.2
38 ± 6
1.1 ± 0.2
< 0.1
48 ± 2
< 0.33
MIGM
(1010 M⊙)
< 0.26
7.4 ± 0.8
< 8.3
6.5 ± 1.0
2.4 ± 0.4
< 1.4
13.8 ± 0.6
< 2.9
†From VLA data (Verdes-Montenegro et al. 2001).
∗Obtained from the assumed σV -- TX relation, equation (3).
aPonman et al. (1996).
bOsmond & Ponman (2004).
cMahdavi et al. (2005).
HCG 15
10-4
HCG 37
10-5
2
c
e
s
c
r
a
/
s
/
s
t
c
)
%
(
.
i
d
s
e
R
40
20
0
-20
-40
10-5
2
c
e
s
c
r
a
/
s
/
s
t
c
10
100
r (arcsec)
HCG 40
)
%
(
.
i
d
s
e
R
30
15
0
-15
-30
10
100
r (arcsec)
2
c
e
s
c
r
a
/
s
/
s
t
c
10-5
)
%
(
.
i
d
s
e
R
60
30
0
-30
-60
2
c
e
s
c
r
a
/
s
/
s
t
c
10-4
10-5
)
%
(
.
i
d
s
e
R
60
30
0
-30
-60
1
10
r (arcsec)
HCG 97
100
10
r (arcsec)
100
Figure 2. 0.3 -- 2 keV exposure-corrected surface brightness profiles of groups with detectable diffuse emission, along with best-fitting β -- models (solid lines).
Horizontal dotted lines mark the estimated background level in each case. For each plot, the bottom panel shows fit residuals relative to the best-fitting model.
2
c
e
s
c
r
a
/
s
/
s
t
c
2
c
e
s
c
r
a
/
s
/
s
t
c
2
c
e
s
c
r
a
/
s
/
s
t
c
2
c
e
s
c
r
a
/
s
/
s
t
c
2
c
e
s
c
r
a
/
s
/
s
t
c
10-6
10-7
10-5
10
10-5
10-6
10-6
10-5
10
HCG 7
10
100
HCG 30
100
HCG 44 (Chandra)
10
100
HCG 44 (XMM)
10
100
HCG 100
100
r (arcsec)
Figure 3. As Fig. 2, but for the groups without detectable diffuse emission.
the upper limits to MIGM for the X-ray undetected groups conser-
vatively assume a uniform IGM distribution. If instead assuming a
standard β -- model for the hot gas in these groups, with central den-
sity equal to the inferred mean value hnei and with typical group
values of, e.g., β = 0.5 and rc = 20 kpc, the derived IGM mass
limits would be reduced by factors of 4 -- 7.
4.1 X-ray properties of the intragroup medium
4.1.1 HCG 7
This group remained X-ray undetected in shallow ROSAT All-Sky
Survey (RASS) data. Despite the XMM data of this target represent-
ing the deepest X-ray observation within our sample, no diffuse X-
ray emission is detected in the group when comparing the emission
level of the exposure-corrected and particle-subtracted 0.3 -- 2 keV
mosaic image inside r = 9 arcmin (r ≈ 140 kpc) with that mea-
sured in a surrounding annulus. This is corroborated by the derived
surface brightness profile shown in Fig. 3. Although this profile
does seem to hint at a weak signal inside r ≈ 2 arcmin, the com-
bined signal inside this region is significant at less than 1.8σ, is
not picked up by 'vtpdetect', and is not discernible in the smoothed
image presented in Fig. 1. Thus, we conservatively treat it as a non-
detection.
Ram pressure stripping in Hickson groups
7
In fact, no extended emission unassociated with individual
galaxies is identified by 'vtpdetect' within the 9 arcmin radius, with
the exception of the source visible in Fig. 1 roughly ∼ 5 arcmin
south of the optical group centre. There are no optical/infra-red
counterparts to this X-ray source listed in NED within a 2 arcmin
diameter, despite the proximity of the group (D ≈ 54 Mpc). The
emission is detected out to 2.2 arcmin from the X-ray centroid at
3σ above the local background, confirming that it is clearly ex-
tended. A thermal plasma model fit to the spectrum extracted from
the pn data within r = 1.5 arcmin of the centroid provides an ac-
ν = 0.92 for 24 degrees of freedom
ceptable fit, with a reduced χ2
(d.o.f.). This yields a best-fitting temperature T = 2.6+0.7
−0.4 keV
and redshift z = 0.41+0.12
−0.02 for an assumed abundance of 0.3 so-
lar. A simple power-law model with Galactic absorption yields
Γ = 1.84 ± 0.10 for the power-law index, but the fit is not pre-
ferred to a thermal model (red. χ2
ν = 1.07 for 25 d.o.f, i.e. a
change of ∆χ2 = 4.8). The estimated temperature can be com-
pared to that expected for the IGM in HCG 7 on the basis of its
very low galaxy velocity dispersion, σV = 95 km s−1, for which
the σV -- TX relation of Osmond & Ponman (2004) would suggest
only T = 0.3 ± 0.1 keV. Combined with the redshift estimate of
z ≈ 0.4, this strongly suggests that this emission is not associ-
ated with HCG 7 itself. The X-ray centroid also coincides to within
10 arcsec with an NVSS source with a 1.4 GHz flux of 19.4 mJy
(corresponding to 1 × 1025 W Hz−1 at z = 0.41), so the X-ray
emission is conceivably associated with a z ≈ 0.4 background
cluster harbouring a central radio-loud galaxy.
For T anywhere in the range 0.2 -- 0.4 keV and assuming any
subsolar metallicity, our failure to detect IGM emission inside r ≈
150 kpc translates into a 3σ upper limit to the unabsorbed 0.3 --
2 keV luminosity inside this region of LX < 5 × 1039 erg s−1,
with a corresponding limit to the mean gas density of hnei < 6 ×
10−5 cm−3. We note that HCG 7 is included in the group catalogue
of Yang et al. (2007), with a total group mass, as estimated from its
optical properties, ranging from 2.4 -- 5.0×1012 M⊙ depending on
the method assumed. This places HCG 7 at the very low-mass end
of the group mass function, with a mass similar to that of the Local
Group. Thus, it is perhaps not surprising that we fail to detect any
IGM emission in this system.
4.1.2 HCG 15
This group has a relatively high velocity dispersion of σV ≈
400 km s−1, and hot intragroup gas was already detected in pointed
ROSAT observations (P96). Significant IGM emission is seen in the
XMM data presented in Fig. 1, revealing a somewhat disturbed X-
ray morphology. Despite this irregularity of the emission, an op-
tically bright early-type galaxy is present roughly at centre of the
X-ray emission, as is typical for fairly undisturbed X-ray bright
groups. The facts that this galaxy is a lenticular rather than an ellip-
tical, and that the IGM emission is not (yet) strongly peaked on this
galaxy, may suggest that the group is in the late stages of dynamical
relaxation.
Emission is detected in the imaging data out to r = 8.9 arcmin
(r ≈ 230 kpc) at 5σ above the X-ray background level evaluated
from a surrounding annulus. A radial surface brightness profile is
shown in Fig. 2, extracted from the centroid of emission in bins
containing a signal-to-noise ratio of S/N > 5. As expected from
the irregular X-ray morphology, a standard β -- model is not a satis-
factory description of this profile, with χ2
ν of 3.5 for 52 d.o.f. The
data show significant deviations from the best-fitting model (with
β = 0.39 ± 0.01 and rc = 21.5+5.5
−4.8 arcsec) at all radii. However,
8
J. Rasmussen et al.
the fit residuals do not exhibit any systematic radial variation, sug-
gesting they are caused by local fluctuations in the IGM distribution
rather than large-scale inhomogeneities. Hence, despite the fact that
the β -- model fit is clearly statistically unacceptable, it remains use-
ful for our purposes as a means of characterizing the global hot gas
distribution. From inspection of Fig. 2, it is also not clear that a dif-
ferent, or more complex, model would be able to provide a better
description.
The relatively broad point spread function (PSF) of XMM, not
accounted for in the surface brightness fitting, could potentially af-
fect the observed profile at small radii, and hence the derived core
radius and central gas density. We do not expect this to be an im-
portant effect, however, because even just the innermost radial bin
in Fig. 2 extends to r = 12 arcsec, roughly twice the full-width at
half maximum of the EPIC PSF. The fact that the best-fitting core
radius is another factor of two larger also suggests that PSF blurring
does not have a significant impact on the derived results.
Using the double-subtraction approach for extracting a back-
ground spectrum, a fit to the global 0.3 -- 5 keV spectrum extracted
inside r = 6 arcmin (r = 150 kpc) gives a temperature T =
0.83+0.09
−0.06 keV and abundance Z = 0.03 ± 0.01 Z⊙, thus confirm-
ing the low abundance derived from ROSAT data within the same
region (Osmond & Ponman 2004). These values are consistent with
those obtained using local background subtraction, but results are
better constrained due to the superior statistics of the blank-sky
background data. The derived flux and surface brightness profile
imply a central hot gas density n0 = 4.4 ± 0.5 × 10−3 cm−3.
4.1.3 HCG 30
Despite this group representing the most HI deficient system within
our sample, the Chandra data do not reveal any clear evidence
for diffuse IGM emission. No extended sources outside individual
galaxies are detected by 'vtpdetect', thus corroborating the RASS-
based result of P96. With a galaxy velocity dispersion of only
72 km s−1, the Osmond & Ponman (2004) scaling relations would
suggest a very low IGM temperature of T = 0.2±0.1 keV. In order
to test for the presence of any such gas, we generated 0.2 -- 0.4 keV
images of the data on the S2 and S3 CCDs separately. These im-
ages were exposure-corrected and smoothed but not background-
subtracted, in an attempt to suppress any bias related to ACIS cal-
ibration uncertainties at these low energies. The results reveal no
clear spatial variations in the diffuse emission on either chip, sug-
gesting emission at a level consistent with the local background.
As a further test, we searched the unsmoothed data for a radial
gradient in the exposure-corrected 0.3 -- 2 keV emission level across
the S2 and S3 CCD's, finding no significant variation with distance
from the optical group centre (see Fig. 3 for a surface brightness
profile extracted on the S3 CCD). This implies that any diffuse IGM
emission would have to be near-uniformly distributed on scales of
∼ 300 kpc, an unlikely scenario for this low-σ system, in which the
angular extent of the region encompassing the four principal group
members is only ≈ 5 arcmin (∼ 90 kpc).
The RASS 0.5 -- 0.9 keV count rate in a 0.5 -- 1 deg annulus cen-
tred on the optical group centre is 8σ above the exposure-weighted
mean of the appropriate Chandra blank-sky data, suggesting a con-
siderable contribution from either background or (Galactic) fore-
ground emission at this position. The high background rate relative
to blank-sky data requires us to evaluate the background from the
source data, thus restricting the source region under investigation
on S3 to within r ∼ 3 arcmin (∼ 50 kpc) of the optical group cen-
tre. Assuming that the background in the data can be safely evalu-
ated from source-free regions on S3 outside this central region (an
assumption supported by Fig. 3), the density of any hot IGM in the
group can be constrained. Inside this region, and for T in the range
0.1 -- 0.3 keV and any subsolar metallicity, the 3σ upper limit to the
mean gas density is hnei < 1.2 × 10−3 cm−3. The possibility
that T is very low in this group propagates into a relatively weak
constraint on hnei.
4.1.4 HCG 37
Irregular diffuse X-ray emission was detected in this group with
ROSAT out to r ∼ 8 arcmin (Mulchaey et al. 2003), well beyond
the region covered by a single ACIS chip in our Chandra data.
The latter clearly indicate that the group emission is sharply peaked
on the early-type galaxy HCG 37a (= NGC 2783), the nucleus of
which is also detected as a point-like source in the data. The as-
sociation of the IGM X-ray peak with HCG 37a was not obvious
from the earlier ROSAT data, as the much broader ROSAT PSF re-
quired Mulchaey et al. (2003) to exclude point-like emission out
to r = 1.5 arcmin from the peak, thus effectively masking out
HCG 37a in the data. Despite the overall irregularity of the group
emission, the X-ray centroid (when masking out the HCG 37a nu-
cleus) coincides to within 10 arcsec with the optical position of
HCG 37a as listed in NED.
Since group emission covers the S3 CCD, we cannot reliably
use any method relying on source-free regions on S3 to evalu-
ate the background in the Chandra data. The situation is aggra-
vated by the fact that RASS data indicate a 3σ soft background
deficit at this position relative to the appropriate blank-sky data,
so background subtraction by means of these is not straightfor-
ward either. To circumvent these issues, we adopted the method
employed by Vikhlinin et al. (2005). First, a source minus blank-
sky spectrum was extracted on the back-illuminated S1 chip, to
quantify the difference in the soft background between source- and
blank-sky data. The spectrum was fitted with a T = 0.18 keV
Z = Z⊙ mekal plasma in the 0.4 -- 1 keV range, with the nor-
malization allowed to be negative. The best-fitting model was then
added to the model fit of the blank-sky subtracted source emission
on S3 inside r = 3.7 arcmin (r ≈ 100 kpc) after scaling to the
source region area. The resulting background level was also used
for the surface brightness analysis. We note that, at 90 per cent
confidence, the best-fitting T and Z resulting from this approach,
T = 0.86+0.13
−0.04 Z⊙, are just consistent
with the P96 values inside r = 150 kpc (T = 0.67 ± 0.11 keV,
Z = 0.17 ± 0.15 Z⊙), lending some credibility to this approach.
−0.09 keV and Z = 0.10+0.08
The surface brightness profile shown in Fig. 2 confirms the
presence of emission across the full S3 CCD in the Chandra data.
The profile was centred on the X-ray peak and extracted in bins
containing at least 30 net counts. Despite the irregularity of the
emission on large scales (Fig. 1), a β -- model provides a good fit
to the profile across the full radial range plotted in Fig. 2, yielding
β = 0.42 ± 0.02 and rc = 3.5+1.5
ν = 0.80 for
19 d.o.f. The spatial and spectral results imply a gas density in the
group core of 0.038 ± 0.006 cm−3.
−1.6 arcsec, with χ2
4.1.5 HCG 40
The Chandra observation of this group was split into two sepa-
rate pointings, so a merged event file was produced for the imaging
analysis. Although undetected in a 3.6-ks ROSAT pointing (P96),
Fig. 1 suggests the presence of diffuse emission in this group. The
centroid of this emission, with the optical extent of the individual
group members masked out, is located ∼ 0.5 arcmin to the NW
of HCG 40c and so is not clearly associated with any individual
galaxy. To test that the emission seen in Fig. 1 is truly extended and
not simply due to the smoothing of point sources, a surface bright-
ness profile of the unsmoothed, exposure-corrected emission was
extracted from the centroid in bins of at least 30 net counts, with
individual galaxies masked out. The result is shown in Fig. 2, with
the background evaluated from off -- source regions on the S3 CCD.
Emission is detected above this background out to r = 2.3 arcmin
(r = 65 kpc), suggesting group-scale extended emission, although
the detection is only significant at > 3σ for the innermost 30 kpc. A
β -- model provides an acceptable fit to this profile, with χ2
ν = 1.01
−0.15 and rc = 46.9+35.8
for 7 d.o.f., yielding β = 0.60+0.40
−18.1 arcsec
(23+17
−9 kpc), in accordance with expectations for a typical X-ray
bright group.
For the spectral analysis of this emission, spectra and response
products were extracted separately for each of the two observations.
The spectra were then jointly fitted within the central r = 1 arcmin,
within which the signal allows useful constraints to be obtained,
using a surrounding 2.5 -- 3.5 arcmin annulus for background esti-
mation. With only ∼ 140 net counts, the IGM abundance remains
unconstrained. Fixing Z at 0.3 solar yields T = 0.59+0.11
−0.12, and T
remains consistent with this for any subsolar Z. For these param-
eters, the observed flux translates into a central electron density of
1.1 ± 0.2 × 10−3 cm−3 and implies a diffuse 0.3 -- 2 keV luminosity
inside r = 2.3 arcmin of 3.1 ± 0.5 × 1040 erg s−1.
The extent of the emission, coupled with the fact that it is
not clearly centred on any group member, suggests that the emis-
sion is not due to, for example, hot gas associated with an el-
liptical but rather reflects the presence of a hot IGM. This in-
terpretation would place HCG 40 among the relatively rare ex-
amples of spiral-dominated groups showing intergalactic hot gas;
within Hickson's (1982) catalogue, only HCG 16, 57, and the well-
studied HCG 92 (Stephan's Quintet) share similar features (e.g.,
Dos Santos & Mamon 1999; Fukazawa et al. 2002; Trinchieri et al.
2003). Based on the B-band luminosities of the group mem-
bers, and on the LX -- LB relations for ellipticals and normal star-
forming spirals from O'Sullivan, Ponman & Collins (2003) and
Read & Ponman (2001) respectively, one would expect a total
galactic diffuse LX ≈ 9 × 1040 erg s−1 in the group, a factor
of three larger than that found here for the intragroup emission. Al-
though care has been taken in masking out emission from the group
members, the low S/N and the compactness of the galaxy config-
uration implies that we cannot exclude a residual contribution to
the diffuse emission from individual galaxies. A conservative ap-
proach would be to regard the association of the observed diffuse
emission with an intragroup medium in HCG 40 as tentative rather
than conclusive.
4.1.6 HCG 44
While Fig. 1 does not indicate the presence of any IGM emission
in this system, this could simply be an artefact of the proximity of
the group (D ≈ 23 Mpc) in combination with the limited Chandra
angular coverage, which furthermore renders quantitative analysis
of the background level in the data non-trivial. Using local back-
ground subtraction could potentially produce unreliable results, as
IGM emission might cover the entire ACIS array. The situation is
further complicated by an enhanced particle level in the cleaned
data compared to the blank-sky files, with the 10 -- 12 keV count
rate on the S3 CCD being 35 per cent higher than the correspond-
Ram pressure stripping in Hickson groups
9
ing blank-sky value. In addition, soft Galactic 0.5 -- 0.9 keV emis-
sion at a level of 2σ above the blank-sky data is also present at
this position, so it is not obvious that blank-sky data would be an
appropriate choice for background estimates.
Fortunately, the presence of an overlapping XMM pointing
allows an independent test for the presence of diffuse emission
in the group. The Chandra and XMM surface brightness profiles
shown in Fig. 3 suggest no detectable IGM emission close to the
optical group centre. Note that the extraction of these profiles ex-
cluded different position angles due to the optical group centre be-
ing close to the southern (northern) edge of the S3 (EPIC) CCDs,
so the profiles extend in largely opposite directions on the sky. The
XMM profile shows no systematic variation out to r = 15 arcmin
(r ≈ 100 kpc), remaining largely consistent with the background
level evaluated outside this region in the source data. Consequently,
the background level for the Chandra profile was estimated from
the northern corners of the S3 CCD, the result suggesting no ex-
cess diffuse emission extending northwards either.
Furthermore, no extended sources that can be unambiguously
associated with group emission were detected by 'vtpdetect'. In ad-
dition to the two group galaxies HCG 44a and b, a third extended
X-ray source is seen on the S3 chip, clearly visible in Fig. 1 roughly
four arcmin north of the spiral HCG 44a, and also seen in both the
XMM data and in pointed ROSAT observations. The X-ray peak of
this source coincides with a 2MASS source with mK = 14.31,
but there is no optical counterpart or redshift information available
in NED. If this source were at the group distance, the resulting
K-band luminosity of 2.1 × 108 L⊙,K would place it at the ex-
treme faint end of the dwarf galaxy luminosity function, with the
ROSAT flux implying a ratio LX/LK ≈ 0.013, two orders of mag-
nitude above typical values seen even for dwarf starburst galaxies
(Rasmussen, Stevens & Ponman 2004). Spectral fit results provide
further support for the idea that this source is unlikely to be associ-
ated with HCG 44. A thermal plasma model fixed at the group red-
shift returns an unacceptable fit (χ2
ν = 1.61), whereas a significant
fit improvement results when leaving z as a free parameter, yield-
−0.39 and z = 0.24+0.09
ing χ2
for an assumed abundance of Z = 0.3 Z⊙.
ν = 1.37 for 6 d.o.f, with T = 1.98+0.67
−0.07
While the Chandra data are useful in terms of investigating
evidence for hot gas being stripped from individual galaxies in this
very nearby system, it is not clear that these data enable significant
improvements on the hot IGM constraints over the existing 4.7-
ks ROSAT pointing (with its much larger field of view enabling a
more reliable background subtraction in this case). Even the XMM
data can only probe emission from a quarter of the volume inside
r = 100 kpc from the optical group centre, due to the latter being
close to the edge of the XMM field of view. Using the adopted σ -- T
relation, which would suggest T = 0.4 ± 0.1 keV, the ROSAT
constraint of P96 on the X-ray luminosity inside r = 150 kpc
(LX < 3.6 × 1040 erg s−1 for our adopted distance) translates into
hnei < 1.0 × 10−4 cm−3 for any subsolar metallicity. The cor-
responding XMM constraint of hnei < 1.4 × 10−4 cm−3 applies
within a volume less than 10 per cent of that probed by ROSAT, so
we have adopted the stronger ROSAT limit in Table 2.
4.1.7 HCG 97
This constitutes the most X-ray luminous system within the sam-
ple. A two-dimensional analysis of a 21-ks XMM observation of
this group was performed by Mahdavi et al. (2005), along with op-
tical spectroscopy identifying 37 members. Their XMM data show
a plume stretching to the southeast, beyond the region covered by
10
J. Rasmussen et al.
Fig. 1. Mahdavi et al. (2005) speculate that this plume could repre-
sent gas stripped from one of the central galaxies, but the XMM data
alone cannot establish this, and the Chandra data on the S2 CCD
cannot improve on the situation beyond confirming the presence
and overall morphology of this feature. One of the spectroscopi-
cally identified member galaxies is located within the plume but is
not itself detected in either the XMM or Chandra data. On the S3
CCD, the emission appears fairly regular in Fig. 1, albeit with the
central X-ray contours somewhat elongated towards the south-east.
If masking out this elongated feature, the centroid of the IGM emis-
sion coincides to within 10 arcsec with the position of the optically
brightest group galaxy, HCG 97a, as listed in NED.
As established already by ROSAT observations
(e,g.,
Mulchaey et al. 2003), diffuse emission in this group extends well
beyond the region covered by the S3 CCD, so blank-sky data were
used to evaluate the background for the Chandra surface bright-
ness analysis (the RASS 0.5 -- 0.9 keV background count rate at this
position is in good agreement with the corresponding exposure-
weighted mean value of the blank-sky data). We note, however, that
the Chandra observation was somewhat affected by background
flares, reducing the useful exposure time from 57.9 to 36.3-ks. As
for HCG 44, the background remains high after cleaning, with the
10 -- 12 keV count rate on S3 again being 35 per cent above the
blank-sky value. While bearing this issue in mind, results indicate
that emission is detected above 5σ significance everywhere on the
S3 chip. A fit to the exposure-corrected 0.3 -- 2 keV surface bright-
ness profile, extracted from the X-ray peak and shown in Fig. 2,
yields β = 0.414 ± 0.004 and rc = 6.6 ± 0.6 arcsec, with
ν = 1.45. However, while Figure 2 and the excellent agreement
χ2
of these results with the best-fitting parameters of Mulchaey et al.
(2003) (who find β = 0.41 ± 0.01 and rc < 0.1 arcmin) sug-
gest that our background estimate is not seriously in error, we will
nevertheless base our normalization of the density profile on the
ROSAT results of Mulchaey et al. (2003), given the concern about
the elevated particle background in the Chandra data. Combin-
ing their spectral results and X-ray luminosity (0.3 -- 2 keV LX =
1.20+0.20
−0.24 × 1042 erg s−1 for our adopted distance) with our sur-
face brightness fit then implies a temperature T = 0.97+0.14
−0.12 keV
and central density of 0.048 ± 0.002 cm−3, which are the values
listed in Table 2.
4.1.8 HCG 100
This is a group in which the HI is clearly being stripped from the
galaxies at present, with much of it pulled into a 100 kpc long
tidal tail extending to the southwest from the optical group centre
(Borthakur et al., in prep.). In addition, the VLA data reveal a strik-
ing HI trail extending to the east away from the group core, pro-
truding from one of the galaxies in the field, Mrk 935. This galaxy
is not included in the original Hickson (1982) catalogue, but is a
group member on the basis of its projected distance from the optical
group centre (6.7 arcmin ∼ 130 kpc) and small radial velocity dif-
ference of ∆v ∼ 250 km s−1 relative to the group mean, as listed
in NED. The associated HI feature may therefore indicate ongoing
stripping as the galaxy falls into the group. HCG 100 thus consti-
tutes an excellent laboratory for the study of the processes whereby
HI is removed from individual galaxies and heated. Unlike the case
for the other Chandra observations presented here, this group was
observed using the ACIS-I array, to allow the observed field to fully
encompass all of the interesting HI features mentioned above.
The Chandra observation was split into two, so the imaging
analysis proceeded as for HCG 40. The combined imaging data,
shown in Fig. 1, and the resulting surface brightness profile in
Fig. 3, do not reveal any clear indications of diffuse emission above
the background level as evaluated outside r = 8 arcmin from the
corners of the ACIS-I array. Although Fig. 3 indicates a mild net
excess in the two innermost bins, the signal within this region is
significant at less than 1.5σ. We also note that no extended sources
are detected outside individual galaxies with 'vtpdetect', and that
the group also remained undetected in RASS data (P96). With the
σ -- T relation suggesting T = 0.3 ± 0.1 keV, the absence of an
IGM detection inside r ≈ 100 kpc (r ≈ 5 arcmin) from the optical
group centre implies a 3σ upper limit on the mean IGM density of
hnei < 3.3 × 10−4 cm−3.
4.2 Individual galaxies
The results presented so far demonstrate the presence of a de-
tectable hot IGM within half of our sample only. However, even in
the absence of such gas, there could still be an intragroup medium
present with temperature or density below our detection limits (in-
cluding any HI already stripped from individual galaxies, as evi-
denced by the GBT detection of intergalactic HI in many of our
groups).
To explore the possibility that group galaxies could be inter-
acting with such a medium, and to search for signs of galaxies be-
ing stripped of any hot gas, we present in Fig. 4 a collage of all
group members which were clearly identified as X-ray extended
sources by our source detection algorithms. These images were
adaptively smoothed following the procedure outlined in Section 3.
Exceptions are HCG 30a and Mrk 935, for which a simple Gaussian
smoothing (with σ = 10 arcsec) was employed due to the very low
S/N. For the groups observed by XMM (HCG 7 and 15), the combi-
nation of group distance and the broader EPIC PSF does not enable
a clear distinction between point-like and diffuse emission, so none
of the relevant galaxies has been included in this figure. HCG 44b,
lying on a chip gap in the Chandra data and not covered by the
overlapping XMM pointing, has also been excluded.
The figure does not reveal any clear evidence for galaxies cur-
rently being stripped of any hot gas. In particular, there are no in-
dications of X-ray tails or bow-shock features indicating interac-
tions with a surrounding medium. Such tails have been observed
extending from galaxies in clusters (e.g., Wang, Owen & Ledlow
2004; Sun & Vikhlinin 2005) but seem to be very rare in groups,
with perhaps NGC 6872 and NGC 2276 the most prominent ex-
amples (Machacek et al. 2005; Rasmussen et al. 2006). There is an
indication of an asymmetric structure in HCG 97d even in the un-
smoothed data, with a hint of a tail pointing south, but the sig-
nal is too weak to exclude contamination by a faint point source.
Mrk 935, with the remarkable HI tail extending to the east, also
presents evidence of some X-ray asymmetry in this direction, but
the S/N is again too low to enable firm conclusions.
In addition to diffuse galactic emission, X-ray point sources
in individual galaxies are detected in all groups, with some of these
sources clearly associated with the galaxy nuclei. The relevance
of this is linked to the possibility that galaxies suffering strong
(tidal) interactions could be showing enhanced nuclear activity, for
example associated with a nuclear starburst or strong AGN accre-
tion fueled by a tidally induced gas inflow. For reference, Table 3
lists detected X-ray sources whose position coincides with the op-
tical centre of individual group members, with luminosities of any
point-like component derived as described in Section 3. Note that
in some cases, such as HCG 97e and most of the XMM sources, we
cannot clearly distinguish between nuclear and galaxy-wide diffuse
HCG 30a
HCG 37a
HCG 40a-d
HCG 44a
A
C
D
B
HCG 44c
HCG 97a,d,e
C
A
E
D
HCG 100a,b
HCG 100: Mrk 935
B
A
Figure 4. Smoothed 0.3 -- 2 keV images of individual group galaxies show-
ing diffuse X-ray emission. A horizontal bar marks a scale of 1 arcmin in
each case.
emission, and the classification of these is followed by a '?' in the
Table 3. For HCG 7, however, the tentative identification of nuclear
X-ray activity in three of the four principal galaxies agrees perfectly
with the Spitzer far-infrared results of Gallagher et al. (2008), sug-
gesting our identification is reasonably robust.
Among the principal members in Hickson's (1982) catalogue,
we detect 22 candidate nuclear X-ray sources in 40 galaxies, with
roughly two-thirds of these falling in the groups with a detectable
IGM. The corresponding nuclear source fractions of 64 ± 17 and
44 ± 16 per cent in the groups with and without detectable hot
gas, respectively, are statistically indistinguishable at the 1σ level.
If instead splitting the sample according to HI deficiency, the cor-
responding fractions are 42 ± 15 (high ∆HI) and 67 ± 18 per cent,
a difference which is only just significant at 1σ. The median nu-
clear X-ray luminosities for the two subsamples are also very sim-
Ram pressure stripping in Hickson groups
11
Table 3. Overview of X-ray sources centred on individual group galaxies
as identified by our detection algorithms. Galaxy morphologies were taken
from NED. Column 4 lists the unabsorbed 0.3 -- 2 keV luminosity of any
nuclear component.
Galaxy
Morph.
Source
HCG 7a
HCG 7b
HCG 7c
HCG 15a
HCG 15b
HCG 15d
HCG 15e
HCG 30a
HCG 30b
HCG 37a
HCG 37b
HCG 40a
HCG 40b
HCG 40c
HCG 40d
HCG 44a
HCG 44b
HCG 44c
HCG 97a
HCG 97b
HCG 97c
HCG 97d
HCG 97e
HCG 100a
HCG 100b
Mrk 935
Sa
SB0
SBc
S0
S0
S0
S0
SB0
SB0/a
S0/E7
Sbc
E
S0
SBb
SBa
Sa
E
SBa
SB0
Sc
Sa
E
S0a
S0/a
S0/a
S?
Nuclear?
Nuclear?
Nuclear
Nuclear?
Nuclear?
Nuclear?
Nuclear?
Diffuse
Nuclear
Diffuse + nuclear
Nuclear
Nuclear
Nuclear
Diffuse?
Diffuse? + nuclear
Diffuse + nuclear
Diffuse
Diffuse
Diffuse + nuclear
Nuclear
Nuclear
Diffuse? + nuclear
Diffuse or nuclear
Diffuse + nuclear
Diffuse + nuclear
Diffuse or nuclear
LX,nucl
(erg s−1)
3.1 ± 0.1 × 1040
3.8 ± 0.4 × 1039
4.9 ± 0.4 × 1039
2.6 ± 0.2 × 1040
2.0 ± 0.2 × 1040
6.7 ± 0.1 × 1041
1.9 ± 0.2 × 1040
--
3.0 ± 0.4 × 1040
1.4 ± 0.2 × 1040
0.8 ± 0.4 × 1039
2.4 ± 0.5 × 1039
1.8 ± 0.4 × 1039
--
9.0 ± 0.9 × 1039
1.8 ± 0.2 × 1039
--
--
7.0 ± 0.8 × 1039
1.8 ± 0.4 × 1039
0.9 ± 0.3 × 1039
8.9 ± 0.9 × 1039
0.9 ± 0.3 × 1039
2.7 ± 0.5 × 1039
1.0 ± 0.3 × 1039
1.6 ± 0.4 × 1039
ilar, 4.4 × 1039 (high ∆HI) and 4.9 × 1039 erg s−1, suggesting
that the above conclusions are not strongly biased by a system-
atic difference in limiting X-ray flux between high -- and low-∆HI
groups. We also note that results from optical spectroscopy indi-
cate that ∼ 40 per cent of the principal members in HCGs in gen-
eral show evidence for AGN activity, with a total of ∼ 70 per cent
showing emission lines from either AGN or star formation activity
(Martinez et al. 2007). Our derived fractions are generally brack-
eted by these values, suggesting that our results give a reasonably
reliable picture of the frequency of nuclear activity within our sam-
ple.
With the limited statistics available, there is thus no strong
evidence from the X-ray data alone for enhanced nuclear activity
within a certain kind of groups in our sample. Specifically, if inter-
preting the X-ray bright or highly HI deficient systems as dynami-
cally more evolved, we find no clear indication that the frequency or
strength of nuclear X-ray activity depends on the dynamical status
of the group. However, we note that this result applies to a small
sample and to the principal members only; a complete census of
galaxy membership from optical spectroscopy would be required
to extend this analysis to optically fainter group members and place
this conclusion on a more robust basis. The tentative lack of a clear
enhancement in nuclear X-ray activity among the most HI defi-
cient groups within our sample may tie in with the observation that
star formation activity is not globally enhanced in HCG galaxies
compared to isolated ones (Verdes-Montenegro et al. 1998), as dis-
cussed in more detail in Section 6.
12
J. Rasmussen et al.
4.3 HI deficiency and IGM properties
The observed diversity in the diffuse X-ray properties of these
highly HI deficient groups immediately suggests that galaxy -- IGM
interactions are not the dominant mechanism for driving cold gas
out of the galaxies within our sample and establishing the observed
HI deficiencies. Of course, this conclusion neglects the fact that we
are not uniformly sensitive to the presence of hot gas in the differ-
ent groups. A quantitative comparison of observed HI deficiencies
and derived IGM properties is therefore presented in Figure 5. The
left panel shows ∆HI and hot IGM mass as listed in Table 2, with
both quantities derived within the same region (inside rHI in the
Table). Even when considering the X-ray detected systems alone,
the obvious lack of a positive correlation between the two quanti-
ties immediately suggests that the amount of hot gas in the group
core is not a pivotal factor for HI removal. We note that an identical
conclusion is reached if replacing ∆HI with 'missing' HI mass in
the plot. The strong 3σ upper limit (M < 2.6 × 109 M⊙) on the
IGM mass in the highly HI deficient HCG 7 (with the largest 'miss-
ing' HI mass in the sample of ∼ 1.8 × 1010 M⊙) only reinforces
this conclusion. Also note, as pointed out in Section 4.1, that more
realistic assumptions about the IGM distribution in the X-ray unde-
tected groups could reduce their upper limits to MIGM by perhaps
an order of magnitude, but that this has no bearing on the above
conclusions.
Hence, neither is it surprising that ∆HI does not show a
clear dependence on the characteristic IGM ram pressure plotted
in Fig. 5b and evaluated as the product of σ2
V and the volume-
weighted mean IGM density MIGM/V within the volume V =
HI covered by the radio data, with all quantities taken from
(4/3)πr3
Table 2. Note that the very compact HCG 40 -- in which we cannot
completely rule out a residual galactic contribution to the derived
IGM mass -- stands out among the X-ray detected groups, with
a characteristic ram pressure 1 -- 2 orders of magnitude below that
seen in the X-ray bright systems. For the X-ray undetected groups,
similar comments apply as for Fig. 5a.
Finally, in Fig. 5c we investigate the dependence of ∆HI
on IGM temperature. Thermal evaporation of galactic HI through
heat conduction from the IGM is expected to proceed at a rate
M ∝ T 5/2 if unsuppressed by, for example, magnetic fields. Given
this strong temperature dependence, the lack of a positive ∆HI -- T
correlation suggests that heat conduction is not an important effect
within our sample. The location of the exceptionally HI deficient
HCG 30 in Fig. 5c would seem to pose a particular challenge for
this mechanism.
Overall, Fig. 5 therefore seems to confirm the notion that
HI deficiency is not tightly linked to the presence or nature of an
IGM in these groups. There are some caveats to this interpreta-
tion though. For example, it is worth emphasizing that it is the four
groups with the highest velocity dispersion that are X-ray detected.
If the remaining groups contain warm (T . 106 K) rather than
hot gas, and thus fall well below the σ -- T relation for X-ray bright
systems, our constraints on the hot gas density could seriously un-
derestimate the true IGM density in these systems. Unfortunately,
this possibility cannot be directly tested with the present data. How-
ever, the fact that our X-ray detected systems scatter fairly tightly
around the Osmond & Ponman (2004) relation, as shown in the in-
set in Fig. 5c, may support our use of this relation for predicting
T also for the undetected systems. Furthermore, results for other
X-ray detected groups suggest, if anything, that the poorest sys-
tems tend to have T on the high side for their velocity dispersion
(Osmond & Ponman 2004) although the situation could, of course,
be different for groups that remain X-ray undetected.
Another concern is that σV and hence Pram in Fig. 5b may
not be robustly determined, being based on just a handful of bright
galaxies in most cases. Large corrections to σV would be needed,
however, in order to affect our overall conclusions. A further point
is that any intergalactic HI already stripped from individual galax-
ies could potentially contribute to the IGM mass and ram pressure.
However, for the X-ray detected systems, the total HI mass inside
the GBT beam is of order 5 per cent of the corresponding hot gas
mass (Borthakur et al., in prep.), suggesting that any cold gas can
be neglected for the present purposes.
Despite the appearance of Fig. 5, it is premature to exclude
the possibility that galaxy -- IGM interactions could play a role for
HI removal in some of our groups. For example, ram pressure strip-
ping is expected to occur when the IGM ram pressure exceeds the
gravitational restoring pressure of the galaxy. The efficiency of this
process therefore depends not only on the properties of the IGM,
but also on those of the individual group galaxies. The process of
viscous stripping (Nulsen 1982) shares similar features, and could
be operating even when ram pressure itself is insufficient to remove
any HI. Finally, ram pressure may also indirectly affect the HI in
the disc. For example, many disc galaxy formation models pre-
dict that massive spirals are surrounded by hot gaseous haloes from
which gas may cool out to provide fuel for ongoing star forma-
tion in the disc (e.g., Toft et al. 2002). The removal of this coronal
gas by external forces could contribute to HI deficiency, if the lim-
ited supply of HI in the disc is consumed by star formation with-
out being replenished from the hot halo (strangulation; see, e.g.,
Kawata & Mulchaey 2008). We next seek to quantify the impor-
tance of these various mechanisms.
5 THE IMPACT OF THE IGM: MODELLING
GALAXY -- IGM INTERACTIONS
In an attempt to constrain the role of galaxy -- IGM interactions for
typical disc galaxies in the individual groups, we constructed sim-
ple analytical models of galaxies moving through the hot intragroup
gas in the gravitational potential of each group. As explained in
Section 4.1, the derived HI deficiency for each group should be
viewed as an average for the group members, since it is often non-
trivial to evaluate observed HI masses for the individual members.
For our modelling purposes, we have therefore adopted a single,
fiducial galaxy model, with overall properties broadly matched to
the fairly well-constrained mean properties of the late-type group
members in our sample.
As described in detail below, the group potential and galaxy
orbits are less well determined for each group. Consequently, we
evolve the adopted galaxy model according to four different as-
sumptions for each group, corresponding to two choices for the
gravitational potential, and two for the galaxy orbits within the cho-
sen potential. The variation among the resulting mass losses from
the galaxy can serve as a means to gauge the uncertainties associ-
ated with these assumptions. We will consider two different strip-
ping processes for the cold gas in the disc, viz. classical ram pres-
sure stripping and turbulent viscous stripping. For the gas in any
hot halo, we only consider ram pressure stripping for simplicity.
Ram pressure stripping in Hickson groups
13
(a)
X-ray detected
X-ray undetected
(b)
(c)
H40
H15
H97
H40
H15
H97
H15
H37
H37
H40
H97
H37
Figure 5. HI deficiency and (a) hot gas mass inside the region used for determining ∆HI, (b) characteristic ram pressure, and (c) hot gas temperature for the
various groups. Empty circles represent groups with no detectable hot gas. Inset in (c) shows velocity dispersion vs. TIGM for the X-ray detected groups, with
the Osmond & Ponman (2004) relation, i.e. equation (3), overplotted as a dashed line.
5.1 Group model and galaxy orbits
Our goal is to evaluate the efficiency of ram pressure stripping
and related processes, without resorting to detailed numerical mod-
elling, which is beyond the scope of this work. For this pur-
pose, we would ideally adopt a single value of the ram pressure
for each group. The simplest approach is to assume the classi-
cal analytical Gunn & Gott (1972) stripping criterion along with
a constant ram pressure equal to its peak value. Hydrodynami-
cal simulations involving a constant ram pressure have shown the
Gunn & Gott (1972) criterion to be remarkably accurate in terms
of predicting the disc stripping radius and the mass of gas lost
(Abadi, Moore & Bower 1999), and it remains a reasonable ap-
proximation even when allowing for orbital variations in ram pres-
sure (J´achym et al. 2007; Roediger & Bruggen 2007).
However, recent hydrodynamical simulations of disc galaxies
moving in radial (J´achym et al. 2007) and two-dimensional orbits
(Roediger & Bruggen 2007) within a non-uniform gas distribution
have revealed an important exception to this rule. If the ram pres-
sure changes faster than the characteristic stripping time-scale, as
will be the case for galaxies moving through a highly concentrated
IGM, the Gunn & Gott (1972) criterion tends to overestimates the
stripping efficiency. Since stripping is not instantaneous, an ISM
element may not always be accelerated to galactic escape velocity
before the peak of the ram pressure is over. In such cases the gas
will eventually re-accrete, a possibility not taken into account by
the Gunn & Gott (1972) criterion. In the present study, these con-
siderations could potentially be relevant for several of our groups,
given the fairly small core radii resulting from the surface bright-
ness fits. Instead of using the peak value of the ram pressure in
our modelling, it therefore seems more sensible to adopt an orbit-
averaged mean ram pressure by evaluating the time spent by the
galaxy at a given velocity and IGM density.
To this end, we first derived total mass profiles Mtot(r) for
each X-ray detected group, assuming a spherically symmetric gas
distribution in hydrostatic equilibrium,
Mtot(< r) =
−kT (r)r
Gµmp (cid:18) d ln n(r)
d ln r
+
d ln T (r)
d ln r (cid:19) .
(4)
Since in general we have neither the statistics nor the spatial cov-
erage to constrain T (r) to large radii, we made two assumptions
about the temperature profile which are likely to bracket the actual
temperature distribution in these somewhat disturbed groups. We
either simply assumed T (r) equal to a constant mean value hT i, or
T (r) = −0.5hT i log(r/r500) + 0.67, appropriate for reasonably
undisturbed groups outside any cool core (Rasmussen & Ponman
2007). In the latter case, r500, the radius enclosing a mean den-
sity of 500 times the critical value ρc, was evaluated iteratively un-
til convergence was reached. For hT i we used the measured value
listed in Table 2.
The resulting mass profiles were characterized analytically by
fitting them with standard 'NFW' models (Navarro, Frenk & White
1997), in which the dark matter distribution is described by
ρ(r) = ρc
δc
(r/rs)(1 + r/rs)2 ,
(5)
where δc is a dimensionless parameter related to total group mass
M200, and rs is a scale radius which reflects the more commonly
used concentration parameter c = r200/rs (see Navarro et al. 1997
for details). With M200, and hence δc, fixed from the measured
(extrapolated) mass profile, equation (5) was fitted to this profile
to derive values of c under the two assumptions about T (r) de-
scribed above. The results, summarized in Table 4, imply typical
values of M200 ≈ 1 − 2 × 1013 M⊙ with c ∼ 5 − 10. For these
group masses, derived concentration parameters are thus in good
agreement with expectations from cosmological N-body simula-
tions (Bullock et al. 2001). Note that the assumption of a declin-
ing temperature profile typically reduces the derived group mass
by ∼ 30 per cent while increasing the halo concentration by a fac-
tor of ∼ 2.
For the galaxy orbital configuration within the derived gravi-
tational NFW potential, we assume two different orbits, both radial
and hence going through the group core. The galaxy is assumed
to be experiencing a face-on IGM encounter in either case. These
maximizing assumptions allow us to estimate how important ram
pressure stripping can ideally be in our groups. The two orbits dif-
fer only in terms of the assumed initial position and velocity of the
galaxy, with the galaxy initially at rest at a small clustercentric ra-
dius in the first scenario, and falling towards the group core from a
large radius and with a high initial velocity in the second. Specifi-
cally, the following two scenarios are considered:
(i) For each group, we determine ¯r, the observed (projected)
mean clustercentric distance of the principal galaxies in each group.
14
J. Rasmussen et al.
Table 4. Results of NFW fits to the derived group mass profiles under the
two assumptions for T (r) described in the text.
T isothermal
Group
HCG 15
HCG 37
HCG 40
HCG 97
r500
(kpc)
r200
(kpc)
M200
(1013 M⊙)
342
362
357
379
541
572
565
500
2.0
2.3
2.2
2.7
c
4.5
4.8
4.0
4.7
T declining
Group
HCG 15
HCG 37
HCG 40
HCG 97
r500
(kpc)
r200
(kpc)
M200
(1013 M⊙)
c
317
334
318
349
474
499
474
523
1.3
1.6
1.3
1.8
10.0
11.0
9.0
10.9
The model galaxy is assumed to be initially at rest at a larger clus-
tercentric distance r0, from which it falls freely towards the group
centre. r0 is chosen such that when the galaxy reaches r = ¯r, it has
attained a velocity corresponding to the observed group velocity
dispersion. Typical values are r0 ∼ 100 kpc and ¯r ∼ 40 kpc.
(ii) The galaxy enters the group halo at r0 = r200 with a radial
velocity vr corresponding to the halo circular velocity at this radius,
vr = (GM/r200)1/2. Typical values are r0 ∼ 500 kpc and vr ∼
400 km s−1.
In both cases we follow the galaxy until it turns around, having
completed one passage through the group core. The first scenario is
chosen in an effort to match the observed average position (modulo
projection effects) and galaxy velocity in each group at present. In
practice, it represents galaxy motion fairly close to the group core,
with a mildly varying ram pressure, and so is somewhat reminis-
cent of the 'classical' ram pressure scenario involving a constant,
high ram pressure. The assumption underlying this orbit is extreme,
however, in the sense that the true clustercentric distances will gen-
erally be larger than the observed (projected) ones, which implies
that galaxies will generally spend a larger fraction of their time at
large distance than implied by this assumption. Therefore we also
consider scenario (ii) as a kind of opposite extreme. In this, galaxies
experience a considerably higher peak ram pressure, but this occurs
only relatively briefly. In the following, these two scenarios will be
referred to as orbit (i) and (ii), respectively.
Although these two orbits do not necessarily encompass the
extreme orbital solutions for the group members, they do represent
two rather different cases, thus offering a handle on the uncertainty
in the predicted mass loss related to orbital assumptions. Further-
more, while completely radial orbits may not be very common, we
note that HI deficient spirals in clusters tend to have more eccentric
orbits than non-deficient ones (Solanes et al. 2001), and that cos-
mological infall along filaments would proceed in fairly eccentric
orbits, thus lending some support to our simplifying assumption.
For a detailed discussion of the impact of orbital parameters on the
stripping efficiency, we refer to Hester (2006).
5.2 Galaxy model
For the galaxy model, needed to estimate the gravitational restor-
ing force and hot halo thermal pressure of a 'typical' late-type
group galaxy within our sample, we follow the general approach
described in Rasmussen et al. (2006) which is repeated here for
completeness. The model consists of a spherical dark matter (DM)
halo with density profile
ρh(r) =
Mh
2π3/2
η
rtr2
h
exp(−r2/r2
t )
(1 + r2/r2
h)
,
a hot gaseous halo of the same form, a spherical bulge with
ρb(r) =
Mb
2πr2
b
1
r(1 + r/rb)3 ,
and exponential stellar and gaseous discs, each of the form
ρd(R, z) =
Md
4πR2
dzd
exp(−R/Rd)sech2(z/zd).
(6)
(7)
(8)
Here Mb, Mh, and Md are the total masses of each component, rb
and rh are the scalelengths of the bulge and halo, respectively, rt is
the DM halo 'truncation' radius, Rd is the cylindrical scalelength
of the disc components and zd the corresponding thickness, and
finally
η = {1 − π1/2q exp(q2)[1 − erf(q)]}−1,
(9)
where q = rh/rt and erf is the error function. With this model, the
restoring gravitational acceleration ∂Φ
∂z (R, z) in the direction z per-
pendicular to the disc can be evaluated analytically for each model
component using the equations of Abadi et al. (1999), to whom we
refer for more details.
In order to constrain model parameters, stellar masses of the
HCG members were evaluated from their B- and K-band mag-
nitudes as listed in NED, following the prescription adopted by
Mannucci et al. (2005). For this purpose, only the principal mem-
bers in our eight groups, as listed by Hickson (1982), with mor-
phological types later than S0 were included, yielding a mean stel-
lar mass of 4.1 × 1010 M⊙. For a subset of these galaxies (8
out of 27), maximum disc rotational velocities, indicative of total
galaxy masses, are also available in the Hyperleda database, with
a mean value of 135 km s−1. We note that, in terms of mean stel-
lar mass, this subset is representative of the full sample, showing
hM∗i = 4.4 × 1010 M⊙. We therefore assume a stellar mass of
4 × 1010 M⊙ for the galaxy model, distributed such as to yield a
bulge-to-disc mass ratio of 1/4, appropriate for an Sb/c spiral. Using
the relation of Haynes & Giovanelli (1984), we further assume an
HI mass of 6.7 × 109 M⊙ to ensure that our model galaxy initially
has a 'normal' HI mass for its sample-averaged blue luminosity of
LB = 1.6 × 1010 L⊙.
The scalelengths of the stellar and gaseous disc components
in the model were chosen to ensure that at least 85 per cent of the
stellar and HI mass resides within the average optical disc radius of
rD ≃ 10 kpc, as derived from the size of the D25 ellipse for each
spiral member in our groups. For simplicity, the gas distribution in
the hot gaseous halo is assumed to follow that of the underlying
dark matter, but with a smaller value of rt corresponding to twice
the 'size' rD of the stellar disc, and with a total mass corresponding
to the HI mass in the disc. The hot halo is assumed to be isothermal
at a temperature T ∼ 0.06 keV, the virial temperature correspond-
ing to the maximum allowed disc rotational velocity in the model,
taken to be ∼ 150 km s−1. Once the baryonic model components
have been specified, the parameters of the DM halo are effectively
Table 5. Adopted parameters of the galaxy model. L is the characteristic
scalelength for each component, i.e. rh and rt for the haloes, rb for the
bulge, and Rd and zd for the disc components.
of radius rD is expected to operate at Reynolds numbers Re & 30,
where
Re = MrD/λ,
(11)
Ram pressure stripping in Hickson groups
15
Component
DM halo
Hot gas halo
Stellar bulge
Stellar disc
Cold gas disc
Mtotal
(1010 M⊙)
L
(kpc)
28
0.7
0.8
3.2
0.67
5, 100
5, 20
0.5
3, 0.25
3, 0.25
set by this maximum allowed rotation velocity, and by the require-
ment that the model baryon fraction match the universal value of
∼ 15 per cent. Table 5 summarises the adopted model galaxy pa-
rameters. Note that the cold gas component in the model refers to
the distribution of HI only, as we do not consider any molecular
gas here. In summary, this model roughly reproduces the average
stellar mass, bulge-to-disc ratio, disc size, and maximum disc rota-
tional velocity seen for the spirals in our groups, with a total baryon
fraction consistent with the universal value, and an initial HI mass
as expected for a non-stripped isolated galaxy with these properties.
5.3 Computing mass losses
Combining equation (5) with the measured density distribution of
intragroup gas, time -- averaged values of IGM density hρi and the
square of the orbital velocity hv2
gali experienced by the galaxy in
its orbit can be evaluated. For the averaging time-scale, we only
consider the segment of the orbit for which the ram pressure ex-
erts a significant influence on the ISM, taken to be from the time at
which the instantaneous ram pressure, if sustained, would remove
at least 1 per cent of the cold gas. This is to avoid the artificial sup-
pression of hρi and hv2i that would otherwise result from including
the time spent by the galaxy at large radius where ram pressure is
completely negligible. Note that the corresponding characteristic
ram pressure is independent of group mass and orbital initial con-
ditions, as it depends only on the assumed galaxy model.
Assuming a face-on IGM encounter, the condition for ram-
pressure stripping is then evaluated as
Σg(cid:16) ∂Φb
+
∂Φh
∂z
+
∂Φg
∂z
(10)
+
∂Φ∗
∂z (cid:17) < hρihv2
gali,
∂z
where ∂Φ
∂z (R, z) is the restoring gravitational acceleration in the
direction z perpendicular to the disc, originating from the stellar
bulge (subscript 'b'), dark matter halo ('h'), gaseous disc ('g') and
stellar disc ('∗'), respectively, and Σg is the surface density of cold
gas. Equation (10) is similar to the classical Gunn & Gott (1972)
stripping criterion, but takes into account the mass distribution in
the galaxy rather than simply assuming a homogeneous disc thin
enough to be described solely by its surface density. Its solution
also provides us with the 'stripping region', the region in the (R,
z) -- plane from which gas is permanently lost by the galaxy.
Apart from conventional ram pressure stripping, transport
processes such as viscous stripping (Nulsen 1982), caused by
Kelvin -- Helmholtz instabilities arising at the ISM -- IGM interface,
could also play a role even when the ram pressure itself is in-
sufficient to remove galactic gas (Quilis, Moore & Bower 2000;
Rasmussen et al. 2006). Turbulent viscous stripping of a gas disc
M is the Mach number of the IGM flow past the galaxy, and λ ∝
T 2n−1 is the ion mean free path in the IGM. The expected mass-
loss rate due to this process (Nulsen 1982),
Mvs ≈ 0.5πr2
(12)
Dρvgal,
scales only linearly with galaxy velocity and so could be important
in a wider range of environments than ram pressure itself. We in-
clude this process in the stripping calculations by evaluating equa-
tions (11) and (12) at each point in the orbit and adding up the
total mass loss. Dealing with disc galaxies rather than spheroids,
we have used only half the Nulsen (1982) mass loss rate in equa-
tion (12) because of the correspondingly smaller galaxy surface
area for a given r = rD. For rD itself, we use the smaller of
the sample-averaged value of rD = 10 kpc and the stripping ra-
dius predicted by equation (10). Note that M 6 1 in equation (11)
even if the galaxy is moving supersonically, as the post-shock IGM
flow past the galaxy will always be subsonic. Also note that the
stripping efficiency of both ram pressure and viscous stripping
should be largely unaffected by the presence of a shock front (see
Rasmussen et al. 2006).
Having dealt with the stripping of cold gas from the disc, we
now turn to the removal of any gas situated in a hot galactic halo.
It seems plausible that, at the very least, the cooling and gradual
inflow of any such gas will be disrupted once the external ram
pressure Pram exceeds its thermal pressure Pth. The simulations
of Mori & Burkert (2000) confirm that this condition provides a
reasonable estimate of the mass of a galaxy that will have its hot
halo completely stripped by ram pressure. For the purpose of also
assessing the importance of strangulation for the model galaxy, we
therefore use the derived values of hρi and hv2i to evaluate the
'strangulation region' where Pram > Pth, and the corresponding
mass of affected coronal gas. The aim here is only to develop a
rough picture of the potential importance of IGM interactions for
strangulation, so the effects of viscous stripping are not considered
for the hot halo.
We readily acknowledge that our modelling approach is infe-
rior to detailed numerical models (e.g., Hester 2006), and hydrody-
namical simulations in particular. It is our hope that one can nev-
ertheless have some confidence in the results, given that we have
employed a reasonably detailed galaxy model and are making some
allowance for orbital variations in ram pressure. The advantage of
the adopted method is that it is easily tailored to the specific condi-
tions in individual groups at little computational cost.
5.4 Results
We re-iterate that we are considering four scenarios for each group,
consisting of two separate assumptions about the galaxy orbit, and
two about the IGM temperature (and hence total mass) distribution
in the groups, as specified in Section 5.1. The results of the HI
stripping calculations for each of these four cases are summarized
in Table 6, which lists the derived r.m.s. orbital velocity vrms, the
orbit-averaged IGM number density, the mass of HI lost due to ram
pressure and viscous stripping, and the total fraction of HI stripped
from the initial model reservoir of 6.7 × 109 M⊙. Recall that vrms
is computed only from the point in the orbit where ram pressure
becomes significant and so depends not only on orbital parameters
but also on the IGM distribution.
16
J. Rasmussen et al.
Table 6. Orbit -- averaged galaxy velocities and IGM densities, along with
predicted HI mass loss ∆M due to ram pressure ('rp') and viscous stripping
('vs') after one passage through the group core for the assumed orbits and
group mass profiles. The final column lists the total fraction f of HI lost.
Orbit (i), T isothermal
vrms
(km s−1)
hni
(cm−3)
∆Mrp
(109 M⊙)
∆Mvs
(109 M⊙)
f
354
342
199
362
6.7 × 10−4
1.5 × 10−3
7.8 × 10−4
2.6 × 10−3
0.9
1.2
0.6
1.6
1.0
1.4
0.2
2.3
0.29
0.39
0.12
0.58
Orbit (i), T declining
vrms
(km s−1)
hni
(cm−3)
∆Mrp
(109 M⊙)
∆Mvs
(109 M⊙)
f
404
402
266
435
7.2 × 10−4
1.9 × 10−3
8.2 × 10−4
2.8 × 10−3
1.0
1.4
0.8
1.8
0.9
1.2
0.2
2.0
0.29
0.39
0.14
0.57
Orbit (ii), T isothermal
vrms
(km s−1)
hni
(cm−3)
∆Mrp
(109 M⊙)
∆Mvs
(109 M⊙)
f
498
541
630
536
1.2 × 10−4
1.5 × 10−4
4.3 × 10−5
2.8 × 10−4
0.6
0.7
0.5
0.9
1.2
1.9
0.3
4.3
0.27
0.39
0.12
0.77
Orbit (ii), T declining
vrms
(km s−1)
hni
(cm−3)
∆Mrp
(109 M⊙)
∆Mvs
(109 M⊙)
f
448
491
554
477
1.2 × 10−4
1.5 × 10−4
4.6 × 10−5
2.8 × 10−4
0.5
0.6
0.4
0.8
1.2
1.9
0.3
4.3
0.26
0.38
0.11
0.76
Group
HCG 15
HCG 37
HCG 40
HCG 97
Group
HCG 15
HCG 37
HCG 40
HCG 97
Group
HCG 15
HCG 37
HCG 40
HCG 97
Group
HCG 15
HCG 37
HCG 40
HCG 97
As can be seen from the Table, the amount of gas lost through
either stripping process is generally an appreciable fraction of the
initial HI mass. Viscous stripping in particular can remove a sub-
stantial fraction of the cold disc gas for all model assumptions.
HCG 40 is the exception, with the relatively tenuous IGM in this
group removing at most 10 -- 15 per cent of the HI. The steeper IGM
density profile in this group also implies that a larger fraction of the
total IGM mass is encountered at high velocity, leading to a higher
∆Mrp/∆Mvs ratio than for the other groups, and a higher value
of vrms in orbit (ii), because the ram pressure becomes significant
closer to the core in this system.
The Table further shows that the main variation in the com-
puted mass loss for a given group and stripping mechanism derives
from the choice of orbit rather than the assumed group mass pro-
file. Orbit (i) is generally more efficient at removing gas through
ram pressure stripping than orbit (ii), because although the peak
ram pressure is considerably higher in the latter case, the galaxy
spends a comparatively shorter time in regions corresponding to
high values of Pram. Conversely, viscous stripping is more efficient
in orbit (ii), as this mechanism acts even at relatively low vgal, al-
lowing the associated mass loss to build up significantly over the
much longer crossing time-scale relevant for this orbit. Note, how-
10.0
)
c
p
k
(
z
1.0
0.1
1
HCG 40
HCG 15
HCG 37
HCG 97
R (kpc)
10
Figure 6. Isobars of gravitational restoring pressure (dashed) and hot halo
thermal pressure (dotted) for our fiducial model galaxy in the different
groups. Dashed lines outline the galactic regions outside which HI in the
disc can be stripped by ram pressure, corresponding to equality in equa-
tion (10). Dotted lines show the corresponding regions for the hot halo gas.
The order of the contours is the same in both cases.
ever, that for a given group, the outcome in terms of total HI mass
loss, ∆Mtot = ∆Mrp +∆Mvs, is largely insensitive to the various
orbit and mass profile assumptions, despite clear variations in hni
and vrms among the four scenarios. The only exception to this is
the significantly more massive HCG 97, for which the deeper grav-
itational potential and higher IGM mass implies that orbit (ii) is
relatively more efficient at removing galactic gas than for the other
groups.
Figure 6 outlines the stripping region for a model galaxy in
each of the X-ray detected groups, based on solving equation (10).
The figure shows the one of our four scenarios in which the effect of
ram pressure is generally most pronounced, i.e., orbit (i) with T (r)
declining. In the outer disc, the gravitational restoring pressure, and
hence the stripping region for the cold gas, is seen to be nearly in-
dependent of vertical disc height for interesting values of z, and
the gravitational restoring force at a given R peaks well above the
disc. Compared to the 'size' of the stellar disc (rD = 10 kpc), it
is clear that the gas disc becomes mildly truncated by ram pressure
in HCG 97 and 37 but remains largely unaffected in the other two
groups. Note that this truncation reduces the viscous stripping effi-
ciency by reducing the surface area of the gas disc exposed to the
IGM. Viscous stripping is therefore slightly more efficient for low
values of Pram, adding to the explanation of the higher values of
∆Mvs for orbit (ii) in Table 6.
Predicted HI deficiencies corresponding to the mass losses in
Table 6 are compared to the observed values in Fig. 7. Errors on
predicted deficiencies correspond to the full range in predicted HI
mass loss for each group under the various orbit and mass pro-
file assumptions. Shown are the expectations from ram pressure
stripping alone, as well as the full HI mass loss from combining
equations (10) and (12). The dashed line in the figure represents
equality between modelled and observed HI deficiencies; under our
model assumptions, anything below this line cannot be explained
by galaxy -- IGM interactions alone. The plot clearly suggests that
ram pressure stripping on its own is not sufficient to cause the ob-
r.p. + visc. stripping
r.p. stripping only
H97
H37
H37
H15
H97
H15
H40
H40
Figure 7. Observed HI deficiencies compared to the model predictions of
Table 6 for the stripping of HI by ram pressure alone (empty diamonds),
and by ram pressure plus viscous stripping (shaded). Dashed line represents
equality between observed and predicted ∆HI.
Table 7. Fraction of hot halo gas lost due to ram pressure in the various
orbital scenarios.
Group
Orbit (i)
T (r) const.
Orbit (i)
T (r) decl.
Orbit (ii)
T (r) const
Orbit (ii)
T (r) decl.
HCG 15
HCG 37
HCG 40
HCG 97
0.53
0.74
0.31
0.84
0.64
0.85
0.46
0.94
0.31
0.39
0.21
0.52
0.27
0.35
0.18
0.48
served HI deficiencies, even in this X-ray bright subsample of our
groups. When including viscous stripping, however, if as efficient
in removing HI as assumed here, galaxy -- IGM interactions can cer-
tainly help explain observed values of ∆HI, but they can potentially
fully account for the HI loss only in HCG 37 and HCG 97, i.e. in
just two out of our eight groups.
Regarding the issue of strangulation, Fig. 6 also illustrates the
derived stripping region for hot halo gas according to the adopted
Pram > Pth criterion. The figure suggests that ram pressure alone
could remove a sizable fraction of the halo gas in the model. Ta-
ble 7 lists the fractions of stripped halo gas in the different scenar-
ios, showing that these can be substantial for orbit (i) in particular,
peaking at ∼ 95 per cent in HCG 97. Viscous stripping, not in-
cluded here, could potentially contribute beyond these estimates.
This suggests that galaxy -- IGM interactions in X-ray bright groups
could play an important role in removing the gas supply that may
ultimately fuel star formation in spirals, in qualitative agreement
with the simulation results of Kawata & Mulchaey (2008). For our
specific model setup, there is a large dispersion in the fraction of
gas affected, however, and it is not clear that the effect would be
important in groups such as HCG 40.
Ram pressure stripping in Hickson groups
17
5.5 Model limitations and caveats
It is important to stress that the calculations presented here are only
intended to provide a rough picture of the impact of galaxy -- IGM
interactions in our sample, and we do not claim that these results are
anything but indicative. Caveats include the fact that the adopted
galaxy model parameters provide a plausible, but not necessarily
unique, representation of the spirals in our sample. It is also a
possibility, albeit one we cannot easily evaluate, that some group
members may be individually less HI deficient than the 'group-
averaged' value of ∆HI, and so could have HI deficiencies con-
sistent with removal by ICM interactions alone, even if our model
results suggest otherwise for the group members as a whole.
As regards the model calculations, an important limitation is
the fact that the model is completely static. In practice, the gas
distributions both in the groups and their galaxies will be evolv-
ing, which could be particularly relevant for orbit (ii) where orbital
time-scales are several Gyr. Also, the estimate of viscous stripping
mass loss does not take into account that this process itself will re-
duce the size of the gas disc and hence the mass loss rate according
to equation (12). Furthermore, this stripping process may saturate
(Nulsen 1982) which would reduce the mass loss, the presence of
magnetic fields could also suppress hydrodynamical instabilities,
and the presence of a hot halo could perhaps to some extent shield
the cold gas disc from such instabilities. The latter could be of par-
ticular relevance for orbit (ii), in which ∆Mvs is high but where a
comparatively smaller fraction of halo gas is lost due to ram pres-
sure. These considerations suggest that the estimated contribution
by viscous stripping to the HI mass loss should be regarded as an
upper limit.
With their compact galaxy configurations and low velocity
dispersions, our groups could also represent environments in which
galactic dark matter haloes are subject to significant tidal trunca-
tion. Hence, another issue is to what extent our results are affected
by the adopted assumptions on dark matter in the galaxy model.
For the adopted disc and bulge parameters, our freedom to modify
the assumed DM distribution is mainly constrained by the require-
ment that the maximum disc rotational velocity vmax of the model
should not exceed the allowed ∼ 150 km−1. A full exploration of
model parameter space is beyond the scope of this work, and we
refer to Hester (2006) for a more thorough investigation of these is-
sues. Nevertheless, to provide a rough picture of the impact of DM
for our results, we repeated the stripping calculations of Section 5.3
with three different modifications, subject to the above vmax crite-
rion:
(i) Assuming no DM [or larger DM halo scalelengths rt and rh
in equation (6)]. This must be regarded as an extreme assumption,
which should facilitate stripping.
(ii) Increasing the DM mass by a factor of two (for which the vmax
criterion then requires at least one of the DM scalelengths rt and
rh to increase by a similar factor). This should suppress stripping.
(iii) Assuming a higher DM concentration (i.e. lower rh or rt) by
a factor of three (suppresses stripping), which in turn requires a
lower total DM mass by a factor of five (facilitates stripping). These
choices comply with the typical tidal truncation of DM halos in-
ferred for galaxies in massive clusters (Limousin et al. 2007).
We find that, in all three cases, the amount of HI lost by the model
galaxy is at most modified by 10 -- 15 per cent, regardless of the
group and orbit considered. This perhaps somewhat surprising re-
sult has its origin in the fact that the vast majority of cold gas
in the model is located at low galactocentric distances, where the
18
J. Rasmussen et al.
restoring gravitational acceleration -- which determines the strip-
ping region in accordance with equation (10) -- is dominated by
the baryonic disc components and not the dark matter halo. The
variations in the stripped gas mass for the hot halo can be larger,
up to 45 per cent for HCG 40, but they can generally easily be ac-
commodated by the variations associated with the different orbital
assumptions. We therefore conclude that our results are reasonably
robust to changes in the assumed amount and distribution of dark
matter in the galaxy model.
6 DISCUSSION
The groups in our sample are all HI deficient even when account-
ing for any intergalactic HI not clearly associated with individual
galaxies. Removal of HI from the group members is by itself insuf-
ficient to explain this situation, as the removed gas must also be pre-
vented from staying neutral. Thus, unless HI is somehow destroyed
in situ within the galaxies, it must go through a two-stage process
whereby it is first removed from the galaxies and then ionized by
a possibly unrelated mechanism. Here we discuss these different
possibilities in light of our X-ray and modelling results.
6.1 In situ destruction of HI
In this context,
The HI could potentially have been destroyed within the galaxies
themselves, either by evaporation though thermal conduction from
the IGM, or through consumption by star formation, provided there
is no continuous replenishment of cold material. A third possible
mechanism involves heating and possibly ejection from the galax-
ies by starburst winds. While the first possibility, direct heating by
the IGM, seems implausible in light of the absence of a positive cor-
relation between ∆HI and TIGM (cf. our discussion of Fig. 5c), the
two other scenarios deserve some further attention. HI consumption
by star formation could help establish observed deficiencies, but a
prerequisite is that the consumed gas is not simply replenished at
a similar rate, for example through the cooling out of hot, coro-
nal material on to the disc. Our model results for the X-ray bright
groups (Table 7) suggest that such a strangulation scenario could at
least be greatly facilitated by the removal of coronal gas due to ram
pressure, but a more detailed exploration of model parameter space
would be required to assess the general validity of this conclusion.
is instructive to compare the results
from our simple analytical model to the simulation results of
Kawata & Mulchaey (2008). These authors investigate the effi-
ciency of ram pressure stripping and strangulation for a galaxy
in a small galaxy group on the basis of a cosmological smoothed
particle hydrodynamics (SPH) simulation. They do not specifically
consider a 'compact' group, but this has the advantage that their
target galaxy is not subject to noticeable gas loss from tidal interac-
tions (D. Kawata, priv. comm.), so in this sense a comparison to our
model calculations is justified. Their target galaxy has initial prop-
erties broadly similar to those of our galaxy model in terms of stel-
lar, gaseous, and total mass. It is followed for one passage through
the group as it enters a group of virial mass M ≈ 8 × 1012 M⊙
from an initial position roughly corresponding to twice the virial
radius. The galaxy orbit thus shows some similarities to our orbit
(ii), but, with a pericentre at r ∼ 200 kpc, does not take the galaxy
through the very core of the IGM distribution. Kawata & Mulchaey
(2008) find that the resulting ram pressure induces strangulation to
a significant degree. Star formation and thus HI consumption be-
comes mildly enhanced during infall, and the hot halo gas is almost
it
completely removed during the group passage. The combination of
these processes has effectively consumed the HI and quenched star
formation by the time the galaxy re-emerges at the virial radius.
Our model conclusion that ram pressure stripping may
strongly affect any coronal gas in the X-ray bright groups is thus in
encouraging agreement with these simulation results, and it is even
possible that our model underestimates the importance of ram pres-
sure in this context. This is made more relevant still if ram pressure
-- or other galaxy interactions with the group environment -- indi-
rectly accelerate strangulation by enhancing the disc star formation
rate and hence the consumption of HI. This possibility draws ob-
servational support from the strong star formation activity seen in
the ram-pressure affected group spiral NGC 2276 (Rasmussen et al.
2006), as well as from recent hydro-simulations (Kronberger et al.
2008). However, Verdes-Montenegro et al. (1998) noted on the ba-
sis of star formation rates derived from IRAS far-infrared luminosi-
ties that there is no indication of enhanced star formation among
the HCG galaxies compared to the level seen in isolated galaxies.
In fact, current star formation rates in these groups generally seem
too low to explain the missing HI by consumption through star for-
mation. For the spirals in our sample, a mean star formation rate of
M∗ ≈ 1.4 M⊙ yr−1 can be derived for the IRAS-detected galaxies.
However, many of our galaxies remain undetected in one or more
IRAS bands and so have only upper limits to M∗ (if including these
M∗ < 0.9 M⊙ yr−1). At
upper limits in the mean, the result is
these rates, the time-scale for star formation to exhaust an initial HI
supply of 6.7 × 109 M⊙ to current levels is at least 5 Gyr, even if
neglecting the return of unprocessed material to the ISM and any
cosmological accretion of gas by the galaxy.
Thus, while our model calculations and the results of
Kawata & Mulchaey (2008) suggest that hot halo gas can be re-
moved fairly efficiently in several of our groups, potentially inhibit-
ing or at least suppressing any replenishment of disc HI from a hot
halo, strangulation does not seem to have played an important role
in establishing current HI levels within our sample. At the observed
star formation rates, the remaining HI can continue to fuel star for-
mation for many Gyr without drastically reducing the HI supply.
Of course, star formation rates could have been much higher in
the past, which could also indirectly have contributed to exhausting
the gas supply through the ejection of gas from the disc by star-
M∗ in these galaxies has generally
burst winds. However, unless
declined to current levels only fairly recently or has been affected
by the group environment over cosmological time-scales, the impli-
cation seems to be that destruction of neutral hydrogen within the
galaxies themselves cannot explain the HI deficiencies. The pres-
ence of intergalactic HI in some of our groups also shows that this
scenario cannot provide an exhaustive explanation. We are there-
fore compelled to also consider the alternative, externally -- driven
removal and destruction of HI.
6.2 HI removal by external forces
Focussing first on the removal of HI, both galaxy -- IGM and galaxy --
galaxy interactions could be envisaged to play a role. Our results
demonstrate that although earlier studies indicated a link between
significant HI deficiency in Hickson groups and the presence of
hot intragroup gas (Verdes-Montenegro et al. 2001), there is clearly
no one-to-one correspondence between the two. The lack of a
detectable intragroup medium inferred here for half of the eight
most HI deficient compact groups specifically appears to rule out
IGM dynamical interactions as generally dominant for HI removal
within these environments. Ram pressure and viscous stripping
could nevertheless still have played a role for HI removal in our X-
ray detected systems. Our modelling results indicate, however, that
the efficiency of these processes is generally insufficient to fully ac-
count for the HI missing from the individual group members, espe-
cially if our simplistic treatment of viscous stripping overestimates
the predicted HI loss, as hypothesized in Section 5.5.
Overall, the simulations of Kawata & Mulchaey (2008) seem
to support these conclusions. In these, ram pressure affects the
amount of cold gas in the disc to an even lesser degree than in
our model, although this could perhaps be attributed to the fact
that our groups are at least twice as massive, and perhaps also to
the maximizing orbital assumptions adopted in our model. Viscous
stripping does not appear to be important in the simulations, but as
noted by Kawata & Mulchaey (2008), such processes are not nec-
essarily well treated by SPH schemes, so a direct comparison to our
results may not be meaningful.
The limited impact of galaxy -- IGM interactions inferred for
our groups commands an alternative explanation for the HI re-
moval. Tidal stripping constitutes an obvious candidate in these
dense, low-σ environments, particularly in light of the fact that
the four X-ray undetected groups within our sample exhibit the
lowest velocity dispersions and so could be expected to represent
environments where tidal interactions should be most important.
However, even if for now ignoring the problem of subsequently
heating the removed HI, it is not immediately clear that tidal in-
teractions would necessarily result in increased HI deficiency, as
they would also affect the stellar component in the galaxies. If stars
and gas are removed in roughly equal proportion, and if the re-
moved stars become part of any undetected intracluster light (cf.
Gonzalez, Zaritsky & Zabludoff 2007) while the HI remains de-
tectable, this could potentially even result in an HI excess. Recall-
ing that HI deficiency is defined here on the basis of the observed
B-band galaxy luminosity, increasing ∆HI through tidal stripping
therefore requires either the preferential removal of cold gas com-
pared to stars (and its subsequent heating), or that LB is simultane-
ously boosted relative to the HI mass for the non-stripped compo-
nents.
The latter possibility gains support from the observation that
specific star formation rates generally tend to be higher in galaxies
with close neighbours (Li et al. 2008), with enhanced nuclear star
formation plausibly arising as a consequence of tidally induced gas
inflow. It is perhaps curious then, as mentioned above, that there is
no evidence for enhanced star formation in Hickson groups com-
pared to the level in isolated galaxies. This runs contrary to expec-
tations for interacting galaxies, and would seem to argue against
LB being significantly boosted in the HCG members relative to
their remaining stellar or HI mass. The other possibility mentioned
above, that HI is more easily tidally stripped than the stellar com-
ponent, is perhaps more promising. In many of our groups, such as
HCG 100 (cf. Section 4.1.8), significant amounts of intergalactic HI
is detected with the GBT, whereas the stellar components are not
noticeably affected. A possible explanation is that a relatively larger
fraction of HI compared to stars initially resided at large galacto-
centric radii, where tidal removal would be most efficient. The fact
that the HI disc in typical spirals and late-type dwarfs is often more
extended than its optical counterpart, with the radial distribution of
HI declining less steeply with radius than that of the B-band light
(Broeils & Rhee 1997; Swaters et al. 2002), seems to support such
an explanation.
Tidal stripping clearly is taking place in some of our groups,
notably in the X-ray undetected groups HCG 100 and HCG 44. In
the latter, both optical (Fig. 1) and HI data (Borthakur et al., in
Ram pressure stripping in Hickson groups
19
prep.) show strong evidence for the SBc galaxy HCG 44d being
tidally stripped. This may indicate that tidal stripping is the primary
mechanism by which HI is removed from the galaxies in these two
groups. Although it is tempting to extend this conclusion to all the
X-ray undetected groups in our sample, and perhaps even beyond,
the inconspicuous star formation rates in Hickson groups in gen-
eral, and the tentative lack of enhanced nuclear X-ray activity in
the highly HI deficient systems in particular (Section 4.2), may not
argue in favour of such an explanation.
6.3 Destruction of removed HI
As emphasized in the beginning of this Section, HI deficiencies
can only be explained if the hydrogen, once removed from the
group members, is also transformed from its neutral phase. Irre-
spective of the processes accomplishing its removal, a mechanism
must therefore also be invoked for ionizing the HI during or follow-
ing its transfer to intergalactic space. In our X-ray bright groups, a
candidate process is readily available, since any removed HI is ex-
pected to evaporate due to heating by the ambient IGM at a rate
M ∝ T 5/2, almost independently of realistic values of the IGM
density (Spitzer 1962). A detailed comparison of the X-ray and HI
properties of these groups should help test this explanation and will
be presented elsewhere (Verdes-Montenegro et al., in prep.). In the
X-ray undetected groups, where any IGM is expected to be rela-
tively cool (cf. Fig. 5c) if at all present, the picture is less clear-
cut. Even if these groups do contain an IGM, the lower predicted
IGM temperatures in this subsample imply average IGM heating
time-scales an order of magnitude above those in the X-ray bright
systems, casting doubt on whether this mechanism would be suffi-
cient.
If gas is predominantly removed by tidal interactions in the
X-ray undetected groups, it is instead conceivable that some of the
HI has been heated by tidal shocks, although the presence of inter-
galactic HI in some of the groups would imply heating time-scales
well in excess of those associated with the HI removal itself. An-
other possibility is that the column density of any removed HI be-
comes too low for the gas to be self-shielding against ionization by
cosmic UV radiation. If so, much of the undetected hydrogen in
these groups could potentially be in the form of a tenuous, photo-
ionized intergalactic plasma. A quantitative investigation of these
possibilities is the subject of future work, but at present a clear pic-
ture of the fate of the removed HI in the X-ray undetected groups
remains elusive.
7 CONCLUSIONS
Based on a sample of eight Hickson compact groups selected
for their high HI deficiencies, we have used Chandra and XMM-
Newton data to assess the properties of any hot intragroup medium
(IGM) and constrain the role of galaxy -- IGM interactions in remov-
ing HI from the galaxies in these groups. The X-ray analysis re-
veals a detectable IGM in four of the eight groups. We have tenta-
tively identified the detected diffuse emission in HCG 40, a spiral-
dominated group, as associated with an intragroup medium, but
the combination of a low signal-to-noise ratio and an exceptionally
compact galaxy configuration precludes a highly robust conclusion
for this particular system. The remaining three groups are all fairly
X-ray luminous, showing substantial amounts of intergalactic hot
gas with a somewhat disturbed morphology in all cases.
20
J. Rasmussen et al.
The remarkable X-ray diversity seen across the sample imme-
diately suggests that the presence of a significant IGM is not a dom-
inant factor in establishing observed HI deficiencies, despite ear-
lier results indicating such a connection (Verdes-Montenegro et al.
2001). It is particularly notable that some of the most HI deficient
groups show no detectable hot IGM, including HCG 30 which only
contains a few per cent of the expected HI mass for its galaxy con-
tent. A comparison of HI deficiency with either hot IGM mass or
characteristic 'mean' ram pressure confirms the lack of a clear cor-
relation even for the X-ray bright systems, although statistics are
naturally too limited to enable firm conclusions on this basis alone.
The HI deficiency does not seem to depend on IGM temperature ei-
ther, suggesting that heat conduction from the IGM does not play an
important role in destroying galactic HI (although once removed,
this gas is likely to evaporate on fairly short time-scales in the X-
ray bright groups).
From fitting analytical models to the derived mass profiles of
the X-ray detected groups, we have constructed plausible models
of the gravitational potential and associated radial galaxy orbits for
each of these groups. Combined with the inferred IGM distributions
and a numerical model of a late-type galaxy with properties broadly
matching those of our observed spirals, this has enabled estimates
of the importance of ram pressure stripping and viscous stripping in
removing HI from the late-type galaxies in each group. The results
indicate that, even under maximizing assumptions about the galaxy
orbit, ram pressure stripping will remove only small amounts of
cold gas from the group members, peaking at 10 -- 25 per cent in the
X-ray bright HCG 97. We find that viscous stripping is generally
more efficient, with the combination of the two processes capable
of removing more than half of the cold ISM in HCG 97 and po-
tentially fully accounting for the missing HI mass in both HCG 37
and HCG 97. However, the efficiency of viscous stripping is likely
overestimated with our simple analytical approach, and yet these
processes are insufficient in terms of explaining the HI deficiency
of the X-ray detected HCG 15 and HCG 40.
The model results also indicate that ram pressure can effi-
ciently remove a large fraction of any hot galactic halo gas that
may otherwise act as a supply of fresh material for star forma-
tion in the disc. However, even if the gas supply to the disc can
be completely cut off, gas consumption at the typical star forma-
tion rates in the groups would proceed far too slowly to explain
the observed shortfall of HI by itself. Much higher star formation
rates in the recent past are required for this process to have had any
significant impact. This may suggest that the observed HI deficien-
cies are not caused by in situ destruction of HI within the galaxies
themselves. It remains a possibility that even modest star forma-
tion activity could have heated some of the HI and lifted it above
the disc midplane where it would be more susceptible to removal
by ram pressure, similar to the situation proposed for NGC 2276
(Rasmussen et al. 2006). The absence of observational signatures
of this process, e.g., in the form of a hot gas tail extending from
any of the group members, may suggest that such a mechanism is
not generally very important within our sample though.
By the process of elimination, it seems plausible that tidal in-
teractions have played a key role for HI removal in the groups, par-
ticularly in those systems containing no detectable IGM. In order
to explain the HI deficiencies, this scenario would likely require
preferential removal of HI over stars, as perhaps facilitated by the
more extended distribution of cold gas relative to stars in typical
late-type galaxies. While it is perhaps not surprising that tidal in-
teractions are affecting the gas content of galaxies in these compact
groups, the tidal stripping explanation still faces some outstanding
issues. Among these is the expectation that such interactions would
generate enhanced star formation or nuclear activity, but there is
no indication that the X-ray faint or highly HI deficient systems in
our sample show evidence for increased such activity (although the
frequency of nuclear activity in galaxies in compact groups in gen-
eral may be rather high; Martinez et al. 2007). If interpreting the
X-ray bright or highly HI deficient systems within our sample as
dynamically more evolved, we thus find no clear evidence that the
frequency or strength of nuclear X-ray activity in the group mem-
bers depends on the dynamical status of the group. It also remains
unclear whether tidal interactions themselves can destroy the HI
during or following its removal from galactic discs and so fully ex-
plain the HI deficiency in any of our groups.
In closing, our results suggest that galaxy -- IGM interactions
can have played a role for the removal and destruction of HI in
some of our groups, but a complete understanding of the origin
of the observed HI deficiencies and the processes causing it is
still lacking. Strangulation or thermal evaporation do not emerge
as important contenders, and typical indirect signatures of tidal
interactions, such as enhanced star formation or nuclear X-ray
activity, are not more pronounced within the more HI deficient
half of our sample. The latter seems in line with previous results
(Verdes-Montenegro et al. 1998) which indicate that star formation
rates in Hickson compact groups are not globally enhanced relative
to the field. We note here that this result could potentially be mis-
leading, however, perhaps masking an evolutionary trend in which
galaxies initially experience enhanced star formation which is then
followed by an environment -- driven suppression. A detailed corre-
lation of HI and X-ray morphology in the groups, coupled with a
broad comparison of individual galaxy properties such as specific
star formation rates, may therefore shed further light on the fate
of the missing HI in these compact systems. This is the subject of
future work.
ACKNOWLEDGMENTS
We thank the referee for useful comments which helped to clar-
ify the presentation of our results. This work made use of the
NASA/IPAC Extragalactic Database (NED) and the Two Micron
All Sky Survey (2MASS) database. Support for this work was
provided by the National Aeronautics and Space Administration
through Chandra Postdoctoral Fellowship Award Number PF7-
80050 and Chandra Award Number GO5-6127X and GO6-7128X
issued by the Chandra X-ray Observatory Center, which is operated
by the Smithsonian Astrophysical Observatory for and on behalf
of the National Aeronautics and Space Administration under con-
tract NAS8-03060. LVM is partially supported by DGI Grant AYA
2005-07516-C02-01 and Junta de Andaluc´ıa (Spain).
REFERENCES
Abadi M. G., Moore B., Bower R. G., 1999, MNRAS, 308, 947
Arnaud M., et al., 2002, A&A, 390, 27
Broeils A. H., Rhee M.-H., 1997, A&A, 324, 877
Bullock J. S., Kolatt T. S., Sigad Y., Somerville R. S., Kravtsov
A. V., Klypin A. A., Primack J. R., Dekel A., 2001, MNRAS,
321, 559
Cortese L., Gavazzi G., Boselli A., Franzetti P., Kennicutt R.C.,
O'Neil K., Sakai S., 2006, A&A, 453, 847
Coziol R., Plauchu-Frayn I., 2007, AJ, 133, 2630
Ram pressure stripping in Hickson groups
21
Trinchieri G., Sulentic J., Breitschwerdt D., Pietsch W., 2003,
A&A, 401, 173
Verdes-Montenegro L., Yun M.S., Perea J., del Olmo A., Ho
P.T.P., 1998, ApJ, 497, 89
Verdes-Montenegro L., Yun M.S., Williams B.A., Huchtmeier
W.K., Del Olmo A., Perea J., 2001, A&A, 377, 812
Vikhlinin A., Markevitch M., Murray S. S., Jones C., Forman W.,
Van Speybroeck L., 2005, ApJ, 628, 655
Wang Q. D., Owen F., Ledlow M., 2004, ApJ, 611, 821
Williams B.A., Rood H.J., 1987, ApJS, 63, 265
Yang X., Mo H. J., van den Bosch F. C., Pasquali A., Li C., Barden
M., 2007, ApJ, 671, 153
This paper has been typeset from a TEX/ LATEX file prepared by the
author.
Dickey J.M., Lockman F.J., 1990, ARA&A, 28, 215
Dos Santos S., Mamon G. A., 1999, A&A, 352, 1
Fukazawa Y., Kawano N., Ohta A., Mizusawa H. 2002, PASJ, 54,
527
Gallagher S. C., Johnson K. E., Hornschemeier A. E., Charlton
J. C., Hibbard J. E., 2008, ApJ, 673, 730
Giovanelli R., Haynes M.P., 1985, ApJ, 292, 404
Gonzalez A. H., Zaritsky D., Zabludoff A. I., 2007, ApJ, 666, 14
Grevesse N., Sauval A.J., 1998, Space Sci. Rev., 85, 161
Gunn J.E., Gott J.R. III, 1972, ApJ, 176, 1
Haynes M.P., Giovanelli R., 1984, AJ, 89, 758
Helsdon S.F., Ponman T.J., 2003, MNRAS, 339, L29
Hester J. A., 2006, ApJ, 647, 910
Hickson P., 1982, ApJ, 255, 382
J´achym P., Palous J., Koppen J., Combes F., 2007, A&A, 472, 5
Kawata D., Mulchaey J.S., ApJL, 672, L103
Kronberger T., Kapferer W., Ferrari C., Unterguggenberger S.,
Schindler S., 2008, A&A, in press (arXiv:0801.3759)
Lewis I., et al., 2002, MNRAS, 334, 673
Li C., Kauffmann G., Heckman T., Jing Y. P., White S. D. M.,
2008, MNRAS in press (arXiv:0711.3792)
Machacek M. E., Nulsen P., Stirbat L., Jones C., Forman W. R.,
2005, ApJ, 630, 280
Limousin M., Kneib J.P., Bardeau S., Natarajan P., Czoske O.,
Smail I., Ebeling, H., Smith G.P., 2007, A&A, 461, 881
Mahdavi A., Finoguenov A., Bohringer H., Geller M.J., Henry
J.P., 2005, ApJ, 622, 187
Mannucci F., Della Valle M., Panagia N., Cappellaro E., Cresci
G., Maiolino R., Petrosian A., Turatto M., 2005, A&A, 433, 807
Martinez M. A., Del Olmo A., Perea J., Coziol R., 2007, in Sa-
viane I., Ivanov V.D., Borissova J., eds, Groups of Galaxies in
the Nearby Universe, Springer-Verlag, Berlin, p. 163
Mori M., Burkert A., 2000, ApJ, 538, 559
Mulchaey J.S., Davis D.S., Mushotzky R.F., Burstein D., 2003,
ApJS, 145, 39
Navarro J. F., Frenk C. S., White, S. D. M., 1997, ApJ, 490, 493
Nulsen P.E.J., 1982, MNRAS, 198, 1007
Osmond J. P. F., Ponman T. J., 2004, MNRAS, 350, 1511
O'Sullivan E., Ponman T. J., Collins R. S., 2003, MNRAS, 340,
1375
Ponman T.J., Bourner P.D.J., Ebeling H., Bohringer H., 1996,
MNRAS, 283, 690 (P96)
Quilis V., Moore B., Bower R., 2000, Science, 288, 1617
Rasmussen J., Ponman T. J., 2007, MNRAS, 380, 1554
Rasmussen J., Ponman T.J., Mulchaey J.S., 2006, MNRAS, 370,
453
Rasmussen J., Stevens I.R., Ponman T.J., 2004, MNRAS, 354,
259
Read A.M., Ponman T.J., 2001, MNRAS, 328, 127
Read A.M., Ponman T.J., 2003, A&A, 409, 395
Roediger E., Bruggen M., 2007, MNRAS, 380, 1399
Sengupta C., Balasubramanyam R., 2006, MNRAS, 369, 360
Solanes J.M., Manrique A., Garc´ıa-G´omez C., Gonz´alez-Casado
G., Giovanelli R., Haynes M.P., 2001, ApJ, 548, 97
Spitzer L., 1962, Physics of Fully Ionized Gases, Interscience,
New York, NY
Sun M., Vikhlinin A., 2005, ApJ, 621, 718
Sutherland R.S., Dopita M.A., 1993, ApJS, 88, 253
Swaters R. A., van Albada T. S., van der Hulst J. M., Sancisi R.,
2002, A&A, 390, 829
Toft S., Rasmussen J., Sommer-Larsen J., Pedersen K., 2002, MN-
RAS, 335, 799
|
astro-ph/0503650 | 1 | 0503 | 2005-03-30T12:11:49 | Star-Cluster Astrometry with Ground-Based Wide Field Imagers | [
"astro-ph"
] | We show the astrometric potential of the Wide Field Imager at the focus of the MPI-ESO 2.2m Telescope. Currently, we are able to measure the position of a well-exposed star with a precision of $\sim$4 mas/frame in each coordinate (under 0.8 arcsec seeing conditions). We present some preliminary results here. | astro-ph | astro-ph |
Astrometry in the Age of the Next Generation of Large Telescopes
ASP Conference Series, Vol. 000, 2005
P. Kenneth Seidelmann and Alice K. B. Monet
Star-Cluster Astrometry with Ground-Based Wide Field
Imagers
Luigi R. Bedin
ESO - Garching, Germany, EU;
[email protected]
Jay Anderson
Dept. of Physics and Astronomy, Rice Univ., Houston, TX, USA;
[email protected]
Giampaolo Piotto, Yazan Momany, Ramakant S. Yadav
Dip. di Astronomia, Univ. di Padova, Padova, Italy, EU;
[email protected]
Abstract. We show the astrometric potential of the Wide Field Imager at
the focus of the MPI-ESO 2.2m Telescope. Currently, we are able to measure
the position of a well-exposed star with a precision of ∼4 mas/frame in each
coordinate (under 0.8 arcsec seeing conditions). We present some preliminary
results here.
1.
Introduction
In the last few years, several Wide Field Imagers (WFIs) at the focus of large
ground-based telescopes have become operative (MPI-ESO 2.2m, AAT 4m, CFH
4m), and their number is continuously increasing (LBT 2×8m, VST 2.5m,
UKIRT 3.8m, VISTA 4m, etc.), as well as their field of view. These WFIs
have allowed us to map completely a number of open and globular clusters in
our Galaxy, and to get accurate photometry for large numbers of stars, with the
additional possibility of studying fast-evolving phases of stellar evolution. By
definition, the WFIs allow large radial coverage in a cluster, so that we can study
the radial distribution of stars in different sequences of the color -- magnitude di-
agram (CMD) and of peculiar objects, which allows us to investigate the effect
of the environment on the evolution of the cluster stars. The wide field cover-
age has made the study of tidal tails in open and globular clusters much more
practicable.
Among the most interesting opportunities offered by the WFIs (still largely
unexplored) are in their astrometric performance. Accurate astrometry over
wide fields is important for a number of reasons. First of all, an accuracy of 0.2
arcsec or better is required to point the fibers of multi-fiber spectroscopic facil-
ities (e.g., FLAMES+GIRAFFE/VLT at ESO). However, the most promising
applications lie in the proper-motion measurements of a large number of stars.
WFIs allow astrometric measurements with an accuracy of far better than 0.2
1
2
left,
Figure 1. On the
an image from the WFI/MPI-ESO 2.2m
([email protected]), centered on the Galactic open cluster NGC 2477. On the right,
OMEGACAM, a WFI four times as large (both in field coverage, and in num-
ber of pixels). OMEGACAM will be mounted at the beginning of 2006 at the
focus of the VST-ESO 2.5m telescope. This telescope has been specifically
designed for wide field imaging, and we we expect that OMEGACAM will
provide better astrometry than the [email protected].
arcsec. With a baseline of a few years, images collected with modern WFIs
can provide proper motions more accurate than those obtainable with old plates
with a baseline of several decades. (Note, though, that these plates will still
remain valuable for long-term non-linear astrometry, such as the determination
of the orbit of long-period visual binaries, and of course for long-term variation
in the light curves).
2. Astrometry: The importance of a careful PSF modeling
In the last year, we have started to apply to wide field ground-based images what
we have learned from Hubble Space Telescope (H ST ) (Anderson & King 2000,
2003, Anderson 2004). For accurate astrometry the ability to reproduce the the
core of the Point Spread Function (PSF) is of crucial importance, much more
so than for accurate PSF-fitting photometry. In fact, the PSF core (where the
derivatives of the stellar profile are highest) contains almost all the astrometric
information. The PSF core needs to be carefully modeled, as does its dependence
on the spatial position within the detector. Moreover, the core of the PSF needs
to be represented with adequate sampling. In Figure 2 we show the precision
achieved in measuring stellar positions in images collected at the [email protected],
under 1′′.2 seeing conditions. We verified that a seeing of 0′′.8 improves the
precision by ∼30%.
At the moment our main limitation comes from the geometric distortion of
the focal plane, and we are working on an algorithm to correct for it. Before
we can correct for it, we must of course understand what the nature of the
distortion is, in terms of (1) what order of polynomial best characterizes it, (2)
whether there is any fine-scale component added by the filters or other optical
3
Figure 2.
Errors in measuring star positions vs. instrumental magnitude,
in a sample of [email protected] images collected under seeing conditions of 1′′.2.
The instrumental magnitude is defined as −2.5 log10 of the Digital Numbers
(DN).
elements and (3) whether it changes over time. Even if we cannot perfectly
characterize the distortion, it is still possible to minimize its effect on astrometry
either by taking a set of exposures at a variety of pointing offsets or by doing
transformations in a more local way (Bedin et al. 2003).
3. Example: M4 field decontamination in 2.2 years
Figure 3 shows an example of the proper-motion potential of ground-based wide
field imagers; we present preliminary results on field-star removal in part of
the low-Galactic-latitude globular cluster M4. Observations collected at the
[email protected] in two epochs separated by a time baseline of just 2.2 years already
allow an excellent separation.
The first-epoch data consist of 3×75s + 2×55s B-band images taken on De-
cember 6, 1999, and the second-epoch data consist of 3×180s images in the same
band, and 3×120s V -band images, taken on February 19, 2002. In order to avoid
first-order color terms, we use only the B images to derive the proper motions,
and we use only the cluster MS stars as reference stars in the transformation,
so that stars moving with the cluster will have zero displacement. The top left
panel of Figure 3 shows the vector point diagram of the displacements, in units
of WFI pixels (238 mas/pixel). From high-accuracy astrometric measurements
on H ST data, Bedin et al. (2003) have shown that the average proper-motion
displacement between cluster stars and field objects is ∼17 mas/yr. Since our
astrometric errors increase rapidly toward fainter magnitudes, we consider as
cluster members all the stars with a proper motion which differs by less than 10
mas/yr from the average proper motion of the cluster MS stars.
4
Figure 3.
[email protected] images separated by only 2.2 yr.
Proper-motion measurements in a field of M4, using two sets of
This example must be considered as an application that is still very pre-
liminary, but it shows the importance of high-accuracy astrometry on wide field
ground-based images.
Acknowledgments. We thank Ivan R. King for a careful reading of the
manuscript.
References
Anderson, J., & King, I. R. 2000, PASP, 112, 1360
Anderson, J., & King, I. R. 2003, PASP, 115, 113
Anderson, J. ACS ISR 04-15, 2004
Bedin L. R., Piotto, G., King, I. R., & Anderson, J. 2003, AJ, 126, 247
|
astro-ph/0208168 | 1 | 0208 | 2002-08-07T22:05:29 | Clusters of galaxies in the local universe | [
"astro-ph"
] | We use a matched filter algorithm to find and study clusters in both N-body simulations artificially populated with galaxies and the 2MASS survey. In addition to numerous checks of the matched filter algorithm, we present results on the halo multiplicity function and the cluster number function. For a subset of our identified clusters we have information on X-ray temperatures and luminosities which we cross-correlate with optical richness and galaxy velocity dispersions. With all quantities normalized by the spherical radius corresponding to a mass overdensity of Delta_M=200 or the equivalent galaxy number overdensity of Delta_N=200Omega^{-1}~666, we find that the number of L>L_* galaxies in a cluster of mass M_{200} is log N_{*666} = (1.44 +/- 0.17)+(1.10 +/- 0.09)log(M_{200}h/10^{15}Msun) where the uncertainties are dominated by the scatter created by three choices for relating the observed quantities to the cluster mass. The region inside the virial radius has a K-band cluster mass-to-light ratio of (M/L)_K=(116 +/- 46)h which is essentially independent of cluster mass. Integrating over all clusters more massive than M_{200}=10^{14}Msun/h, the virialized regions of clusters contain ~7% of the local stellar luminosity, quite comparable to the mass fraction in such objects in currently popular LambdaCDM models. | astro-ph | astro-ph |
Draft version October 26, 2018
Preprint typeset using LATEX style emulateapj v. 04/03/99
CLUSTERS OF GALAXIES IN THE LOCAL UNIVERSE
C.S. Kochanek1, Martin White2, J. Huchra1, L. Macri1, T.H. Jarrett3, S.E. Schneider4
& J. Mader5
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138
2 Departments of Physics and Astronomy, University of California, Berkeley, CA 94720
3 Infrared Processing and Analysis Center, MS 100-22, Caltech, Pasadena, CA 91125
4 Department of Astronomy, University of Massachusetts, Amherst, MA, 01003
5 McDonald Observatory Austin, TX, 78712
email: [email protected], [email protected], [email protected],
[email protected], [email protected], [email protected], [email protected]
Draft version October 26, 2018
ABSTRACT
We use a matched filter algorithm to find and study clusters in both N-body simulations artificially
populated with galaxies and the 2MASS survey. In addition to numerous checks of the matched filter
algorithm, we present results on the halo multiplicity function and the cluster number function. For a
subset of our identified clusters we have information on X-ray temperatures and luminosities which we
cross-correlate with optical richness and galaxy velocity dispersions. With all quantities normalized by
the spherical radius corresponding to a mass overdensity of ∆M = 200 or the equivalent galaxy number
overdensity of ∆N = 200Ω−1
M ≃ 666, we find that the number of L > L∗ galaxies in a cluster of mass
M200 is
log N∗666 = (1.44 ± 0.17) + (1.10 ± 0.09) log(M200h/1015M⊙)
where the uncertainties are dominated by the scatter created by three choices for relating the observed
quantities to the cluster mass. The region inside the virial radius has a K-band cluster mass-to-light ratio
of (M/L)K = (116 ± 46)h which is essentially independent of cluster mass. Integrating over all clusters
more massive than M200 = 1014 h−1M⊙, the virialized regions of clusters contain ≃ 7% of the local
stellar luminosity, quite comparable to the mass fraction in such objects in currently popular ΛCDM
models.
Subject headings: cosmology: theory -- large-scale structure of Universe
1.
INTRODUCTION
Clusters of galaxies have become one of our most impor-
tant cosmological probes because they are relatively easy
to discover yet have physical properties and abundances
that are very sensitive to our model for the formation and
evolution of structure in the universe. Of particular inter-
est to us, a cluster sample provides the means to study the
high-mass end of the halo multiplicity function, the aver-
age number of galaxies in a halo of mass M , which provides
important insight into the process of galaxy formation.
In this paper we have two objectives. First, we will
demonstrate the use of a matched filter approach to find-
ing clusters in a redshift survey using both synthetic cata-
logs and a large sample of galaxies from the 2MASS survey
(Skrutskie et al. 1997, Jarrett et al. 2000). With the syn-
thetic data we can test the algorithm and our ability to ex-
tract the input halo multiplicity function from the output
cluster catalog. Second, we will determine the halo mul-
tiplicity function, N (M ), from the 2MASS survey. This
study is the first phase in a bootstrapping process -- based
on a synthetic catalog known to have problems matching
reality in detail we can calibrate an algorithm which when
applied to the real data can supply the parameters for an
improved model of the data. Then the improved model
can be used to improve the algorithm and so on.
In §2 we describe the synthetic and real 2MASS data.
In §3 we describe our version of the matched filter algo-
rithm. In §4 we test the algorithm on the synthetic catalog,
focusing on our ability to determine the halo multiplicity
function. In §5 we apply the algorithm to the 2MASS sam-
ple. Finally, in §6 we discuss the steps which can improve
both the synthetic catalog and the algorithm.
2. DATA: REAL AND SYNTHETIC
We searched for clusters using galaxies in the 2MASS
survey (Skrutskie et al. 1997) and in simulations of the
2MASS survey. We have chosen the 2MASS survey be-
cause of its uniform photometry and large areal coverage,
which makes it ideal for finding the relatively rare rich clus-
ters in the local neighborhood where a wealth of auxiliary
information is available.
2.1. 2MASS Galaxies
The 2MASS survey provides J, H, K photometry of
galaxies over the full sky to a limiting magnitude of
K <∼ 13.75 mag (Jarrett et al. 2000). Based on the
Schlegel, Finkbeiner & Davis (1998) Galactic extinction
model and the available 2MASS catalogs, we selected all
galaxies with Galactic extinction-corrected magnitudes of
K ≤ 12.25 mag1 and Galactic latitude b > 5◦. The final
sample contains a total of 90989 galaxies distributed over
91% of the sky. The redshift measurements are 89% com-
plete for K ≤ 11.25 mag but only 36% complete between
11.25 < K ≤ 12.25. The galaxies with redshifts are not a
random sample of all of the galaxies -- redshifts are more
likely to be of cluster than field galaxies.
1We use the 20 mag/arcsec2 circular isophotal magnitudes
throughout the paper.
1
2
With the data corrected for Galactic extinction, the re-
lation between apparent and absolute magnitudes is
MK = K − 5 log(DL(z)/r0) − k(z) = K − D(z)
(1)
where r0 = 10 pc, DL(z) is the luminosity distance to
redshift z and k(z) = −6 log(1 + z) is the k-correction.
The k-correction is negative, independent of galaxy type,
and valid for z <∼ 0.25 (based on the Worthey (1994)
models). To simplify our later notation, we introduce
D(z) = 5 log(DL(z)/r0) + k(z) as the effective distance
modulus to the galaxy. We use an Ωmat = 1 cosmological
model for the distances and assume a Hubble constant of
H0 = 100h km/s Mpc. For our local sample the particular
cosmological model is unimportant.
Kochanek et al. (2001) derived the K-band luminosity
function for a subset of the 2MASS sample at comparable
magnitudes. The luminosity function is described by a
Schechter function,
φ(M ) =
dn
dM
= 0.4 ln 10 n∗(cid:18) L
L∗(cid:19)1+α
exp(−L/L∗)
(2)
where MK = MK∗ − 2.5 log(L/L∗) and the integrated lu-
minosity function is
Φ(M ) =Z M
−∞
φ(M )dM = n∗Γ[1 + α, L/L∗].
(3)
The parameters for the global luminosity function are n∗ =
(1.16±0.10)×10−2h3/Mpc3, α = −1.09±0.06 and MK∗ =
−23.39 ± 0.05 using the 2MASS 20 mag/arcsec2 isophotal
magnitudes. This luminosity function is consistent with
that derived by Cole et al. (2001). The total luminosity of
a galaxy is approximately 1.20 ± 0.04 times the isophotal
luminosity.
We should keep in mind that the early-type and late-
type galaxies have different luminosity functions, with the
early-type galaxies being brighter but less common than
the late-type galaxies. The algorithm for finding clusters
requires both a field luminosity function φf (M ) and a clus-
ter luminosity function φc(M ) which will be different be-
cause of the morphology-density relation (e.g. Dressler 1980).
In our present study we assume that we can model both
using the global luminosity function, φf (M ) = φc(M ) =
φ(M ). Once we have catalogs of 2MASS galaxies in dif-
ferent environments we can go back and estimate the
environment-dependent luminosity functions. Note that
we only need the shape of the luminosity function for the
clusters -- all expressions using φc(M ) will appear in the
ratio φc(M )/Φc(MK∗), which eliminates any dependence
on n∗.
2.2. Simulations of 2MASS
We generated a simulated 2MASS catalog based on the
approach described in detail in White & Kochanek (2001;
hereafter WK). We used as a starting point the z = 0
output of a 2563 particle N-body simulation of a ΛCDM
'concordance' cosmology whose predicted mass function
agrees well with recent observational estimates (e.g. Pier-
paoli, Scott & White 2001). Halos were identified using
the friends-of-friends (FoF) algorithm with a linking length
b = 0.2 times the mean inter-particle spacing, and galaxies
were 'assigned' to halos based on the multiplicity function.
The average number of galaxies assigned to a cluster halo
with a FoF mass MF oF is
MB (cid:19)0.9# (4)
MR (cid:19) + 0.7max"1,(cid:18) MF oF
NF oF (M ) =(cid:18) MF oF
galaxies where MR = 2.5 × 1012h−1M⊙ and MB =
4.0 × 1012h−1M⊙ based on the halo occupancy models
of Scoccimarro et al. (2001) which we have modified to
better match the observed number densities of galaxies
and the expectation that 2MASS galaxies will trace the
mass more faithfully than a blue selected sample. We
sometimes used the virial mass M200 defined as the mass
interior to a radius r200 within which the mean density
exceeds 200 times the critical density. The virial radius,
which scales as r200 = 1.6(M200h/1015M⊙)1/3h−1Mpc, is
slightly smaller than the average radius of the FoF cluster
with b = 0.2, so that on average M200 = 0.66MF oF (me-
dian M200 = 0.71MF oF ). For massive halos the number of
galaxies actually assigned to a halo is Poisson distributed
for an expectation value of NF oF . Each halo hosts a 'cen-
tral' galaxy which inherits the halo center of mass posi-
tion and velocity. Any additional galaxies are randomly
assigned the positions and velocities of the dark matter
particles identified with the cluster by the FoF algorithm.
In this way the satellite galaxies trace the density and ve-
locity field of the dark matter.
One problem with standard halo multiplicity functions
is that they are one-dimensional functions, N (M ), nor-
malized to match a specific sample (but see Yang, Mo
& van den Bosch 2002 for improvements in this regard).
Even for galaxy mass scales, current models of the halo
multiplicity function lead to too few galaxies to match
the luminosity or velocity functions of galaxies over broad
ranges (see Kochanek 2001; Scoccimarro et al. 2001; Chiu,
Gnedin & Ostriker 2001). For general theoretical use,
we need the luminosity-dependent multiplicity function,
N (> LM ), for the average number of galaxies more lu-
minous than L in a halo of mass M . For clusters, we
might expect a halo multiplicity function to factor as
N∗(M )Φc(> L)/Φc(> L∗) where Φc(> L) is the inte-
grated cluster luminosity function and N∗(M ) is the mass-
dependent number of galaxies more luminous than L∗. In
the language of Yang et al. (2002) this assumption cor-
responds to assuming a weak dependence of their eΦ∗, eL∗
and eα on M in this range.
With our modified multiplicity function, Eq. (4), we
can match the number density of galaxies in the global
Kochanek et al. (2001) K-band luminosity function or the
K-band galaxy number counts if we view these galaxies
as a complete sample truncated at Lmin = 0.01L∗. The
average number of galaxies assigned to a cluster of mass
M at redshift z in a catalog with magnitude limit Klim is
hN (M, z)i = NF oF (M )C(z). The completeness function
is
C(z) =
Γ[1 + α, L(z)/L∗]
Γ[1 + α, Lmin/L∗]
(5)
where L(z) = max(Lmin, Llim(z)) for the limiting luminos-
ity of 2.5 log(Llim(z)/L∗) = MK∗− (Klim −D(z)). For our
parameters, Llim(z) = Lmin at redshift cz = 1350 km/s
and Llim(z) = L∗ at cz = 14000 km/s. The average num-
ber of galaxies with L ≥ L∗ is
N∗F oF (M ) =
NF oF (M )Γ[1 + α, 1]
Γ[1 + α, Lmin/L∗]
= 0.042NF oF (M )
(6)
for our standard parameters. On average, clusters with
M h/M⊙ = 1015, 1014 and 1013 are assigned 500, 53
and 6 galaxies of which 21, 2 and 0.2 are brighter than
L∗. Over the range we actually find clusters, this rela-
tion is well fit by a power-law of the form log N∗666 =
A + B log(M200h/1015M⊙) with A = 1.32 and B = 0.98.
We adopt this form to describe the halo occupancy func-
tion and summarize all of our subsequent estimates for it
in Table 2.
The model catalog contained 100706 galaxies, whose
clustering properties are close to those of the observed pop-
ulation. Unlike WK, where we characterized the observed
properties of the galaxy simply by the accuracy with which
its redshift could be determined, we assigned K-band mag-
nitudes to each galaxy by drawing randomly from the lumi-
nosity function. This process oversimplifies several aspects
of the real survey. First, we did not include the environ-
ment dependence of the luminosity function. Second, for
lower mass halos where there is essentially one galaxy per
halo, we did not correlate the luminosity with the mass
of the halo. Third, to mimic the redshift measurements
in the 2MASS sample we assumed that all redshifts were
measured for K≤ 11.25 mag and that one-third, randomly
selected, were measured between 11.25 < K ≤ 12.25. This
mimics the statistics of the real data, but the redshift mea-
surements in the real data are more likely to be of clus-
ter galaxies than of field galaxies. Finally we ignored the
b > 5◦ latitude cut made for the 2MASS data and simply
analyzed the synthetic catalogs for the whole sky. Most of
these simplifications should make it more difficult to iden-
tify clusters in the synthetic catalog than in the real data.
Indeed, the visual impression when comparing slices of the
model survey and 2MASS data is that the "fingers of god"
are weaker in the model. We intend to significantly im-
prove our mock galaxy catalogs in the next iteration, using
the relations we derive from the data in later sections (see
§6) combined with a better underlying simulation.
3. FINDING AND ANALYZING THE CLUSTERS
Large cluster samples have been constructed using five
In the first method, used to con-
general approaches.
struct the original
large catalogs and their successors
(e.g. Abell 1958; Shectman 1985; Dalton et al. 1992; Lums-
den et al. 1996; Ostrander et al. 1998; Scoddeggio et
al. 1999; Gladders & Yee 2000; White et al. 1999), clusters
were selected as overdensities in the projected density of
galaxies on the sky. It was quickly realized such surveys
suffer from projection effects and there is a large scatter
between optical richness and cluster mass (for recent theo-
retical studies see e.g. van Haarlem, Frenk & White 1997;
Reblinsky & Bartelmann 1999; WK).
Recently, more quantitative versions of these methods
based on matched filter algorithms to find either the dis-
crete cluster galaxies (Postman et al. 1996; Kepner et
al. 1999; Kim et al. 2001) or the excess luminosity from
unresolved galaxies (Dalcanton 1996; Zaritsky et al. 1997;
Gonzalez et al. 2001) have been developed to find clus-
3
ters at intermediate redshifts. A host of approaches have
recently been applied to the data from the SDSS (see Bah-
call et al. 2002 for recent discussion and Goto et al. 2001
and Nichol et al. 2000 for early work).
Second, with the advent of large redshift surveys, clus-
ter catalogs were constructed by finding three-dimensional
overdensities in the galaxy distribution using the friends-
of-friends (FoF) algorithm (e.g. Huchra & Geller 1982;
Geller & Huchra 1983; Ramella et al. 1994; Ramella,
Pisani & Geller 1997; Christlein 2000, Ramella et al. 2002).
Studies of the FoF algorithm using N-body simulations
(Nolthenius & White 1987; Frederic 1995; Diaferio et
al. 1999) demonstrate the FoF algorithm works well for
appropriate choices of the linking parameters. The tree or
"dendogram" methods (e.g. Tully 1987) are related to FoF
methods.
The third approach to building cluster catalogs is to use
X-ray surveys (Gioia et al. 1990; Edge et al. 1990; Henry
& Arnaud 1991; Rosati et al. 1995; Jones et al. 1998; Ebel-
ing et al. 1998; Vikhlinin et al. 1998; Romer et al. 2000;
Henry 2000; Blanchard et al. 2000; Scharf et al. 2000;
Bohringer et al. 2001; Ebeling, Edge & Henry 2001; Gioia
et al. 2001). X-ray surveys avoid2 many of the problems
in selection and characterization which have made optical
catalogs difficult to use, but frequently lack the sensitivity
to detect and characterize groups or clusters at z ∼> 0.5.
Finally, two rapidly developing approaches are to use sur-
veys based on the Sunyaev-Zel'dovich effect (Carlstrom et
al. 2000) or on weak gravitational lensing (Schneider &
Kneib 1998; Wu et al. 1998; Hattori et al. 1999; Wittman
et al. 2001; Moller et al. 2001; but see White, van Waer-
beke & Mackey 2002).
3.1. The matched filter algorithm
We identify groups and clusters in the 2MASS catalog
using the modified optimal filter method we developed in
WK, which is itself derived from the similar algorithms of
Kepner et al. (1999) and Postman et al. (1996). These
methods identify clusters as 2- or 2.5-dimensional over-
densities in three-dimensional redshift surveys. This is a
departure from the essentially uniform use of the FoF algo-
rithm to find clusters in redshift surveys. While the FoF al-
gorithm clearly works to identify groups and clusters with
very small numbers of galaxies, a matched filter approach
provides several advantages. Matched filters provide like-
lihood estimates for the detection, membership probabil-
ities for individual galaxies, a range of cluster properties
and their uncertainties and (importantly for us) can natu-
rally incorporate galaxies both with and without redshift
measurements.
In WK we used N-body simulations to explore the selec-
tion of clusters in a series of model surveys for clusters at
intermediate redshift. In particular, we developed strate-
gies for selecting samples above a fixed mass threshold over
a range of redshifts, showing that it was relatively easy to
obtain highly complete samples at the price of significant
contamination from real but somewhat less massive clus-
ters. Contamination by detections of non-existent clusters
was rare.
The algorithm works as follows. We model the universe
as a uniform background and a distribution of k = 1 ··· nc
2But see Lewis et al. (2001).
4
clusters. Cluster k is described by its angular position
~θk, (proper) scale length, rck, a fixed concentration, c,
redshift zk and galaxy number Nk. To ease comparisons to
theoretical models we describe the galaxy density profile
by an NFW (Navarro, Frenk & White 1996) profile.
In
three dimensions, the galaxy distribution normalized by
the number of galaxies inside an outer radius rout = crc is
ρ(r) =
1
4πr3
c F (c)
1
x(1 + x)2
(7)
where x = r/rc, F (x) = ln(1 + x) − x/(1 + x) and
r2drρ ≡ 1. The relative number of galaxies in-
side radius r is N (< r) = F (r/rc)/F (c). The projected
surface density is more complicated,
4πR crc
0
Σ(R) =
f (R/rc)
2πr2
c F (c)
(8)
where
f (x) =
1
x2 − 1"1 −
2
x2 − 11/2 tann−1(cid:12)(cid:12)(cid:12)(cid:12)
x − 1
x + 1(cid:12)(cid:12)(cid:12)(cid:12)
1/2# (9)
and tann−1(x) = tan−1(x) (tanh−1(x)) for x > 1 (x < 1).
The relative number of galaxies projected inside cylindrical
radius R is N (< R) = g(R/rc)/F (c) where
g(x) = ln(x/2) +
2
x2 − 11/2 tann−1(cid:12)(cid:12)(cid:12)(cid:12)
x − 1
x + 1(cid:12)(cid:12)(cid:12)(cid:12)
1/2
(10)
(Bartelmann 1996). The projected quantities are consid-
erably more complicated than the simple model (Σ ∝
(1 + x)−2) we used in WK, but the earlier model corre-
sponds to a three-dimensional density which is too compli-
cated for convenient use. We fixed the halo concentration
to c = 4 and the break radius to rc = 200h−1 kpc, as typ-
ical parameters for cluster-mass halos. We estimate the
probability a survey galaxy has its observed properties as
either a field or a cluster member galaxy. These properties
can include the flux, redshift, angular position and color,
with a basic distinction being made between galaxies with
redshifts and those without.
3.2. Probabilities for Field Galaxies
The probability of finding a field galaxy of magnitude
Ki and redshift zi is
Pf (i) = 0.4 ln 10 D2
C (zi)
dDC
dz
φf (Ki − D(zi))
(11)
where DC is the comoving distance. The probability of
finding a field galaxy of magnitude Ki is
Pf (i) =
dN
dm
(Ki)
(12)
where dN/dm =R ∞
ber counts.
0 dzPf (K, z) are the differential num-
3.3. Probabilities for Cluster Galaxies
The probability of finding a cluster galaxy of magnitude
Ki and redshift zi at an angular separation θik from a clus-
ter at redshift zck with (proper) scale length rck, galaxy
number N∗ck, and rest frame velocity dispersion σck is
Pc(i, k) = N∗ck
φc(Ki − D(zck))
rck
Φc(MK∗)
ck(1 + zck)2)
exp(−c2(zi − zck)2/2σ2
Σ(cid:18) DA(zck)θik
√2πσck(1 + zck)
(cid:19)
. (13)
The number of galaxies N∗ck is normalized to represent
the number of galaxies brighter than L∗ inside the ra-
dius rout ≡ crc. This normalization is convenient for
calculation, and we discuss other normalizations in §3.6.
The performance of the algorithm in correctly determin-
ing galaxy membership and cluster velocity dispersions is
significantly enhanced if we optimize the width of the ve-
locity filter for each cluster candidate. We used the range
150km s−1 ≤ σck ≤ 1200km s−1. For a galaxy with an
unknown redshift, the likelihood is
φc(Ki − D(zck))
Pc(i, k) = N∗ck
(14)
Σ(cid:18) DA(zck)θik
rck
(cid:19) .
Φc(MK∗)
We calculate the likelihood in 1h−1 Mpc regions around
each galaxy based on either the measured redshift or the
average redshift for galaxies with the observed magnitude.
The search area was limited to ∆θ = 4◦. The fraction of
cluster galaxies inside search radius R = DA(zck)∆θ of a
cluster at redshift zck is
Ak(∆θ) =
g(R/rc)
F (c)
Φc(Klim − D(zk))
Φc(MK∗)
.
(15)
3.4. Likelihood Function
The likelihood function for finding the first cluster nc =
k = 1 is
∆ lnL(k) = −N∗ckAk +
ln(cid:18) Pf (j) + N∗ckPck(j, k)
Pf (j)
(cid:19)
NgXj=1
(16)
where the sum extends over all galaxies within angle ∆θ of
the cluster. We compute the likelihood at the positions of
each survey galaxy over a range of trial redshifts, optimiz-
ing the number of galaxies N∗ck for each trial. We then add
the trial cluster which produced the largest change in the
likelihood to the cluster sample. We automate the process
and avoid finding the same cluster by including each new
cluster in the density model used to find the next cluster,
∆ lnL(nc) = −N∗cncAnc
+PNg
j=1 ln(cid:18) Pf (j)+Pnc
Pf (j)+Pnc−1
k=1
k=1
N∗ckPck(j,k)
N∗ckPck(j,k)(cid:19) .
(17)
The clusters k = 1 ··· nc − 1 are our current catalog and
have fixed properties, while k = nc is our new trial cluster.
By this method we "clean" the clusters out of the sample
in order of their likelihood (for further discussion see WK).
We found that several small modifications improved the
performance of the algorithm. In most cases these are pri-
ors on the filter variables N∗c and σck designed to stabilize
0 + N 2
parameter estimation for systems with very few galaxies.
First, we added a prior to the likelihood, Eq. (17), of the
form − ln(N 2
∗c) with N0 = 0.1 comparable to the
value of N∗c expected for a very poor group. This en-
codes the information that massive clusters are rare, with
a slope roughly matching the power-law slope of the halo
mass function.
In some senses the prior acts to control
the problem of Malmquist biases created by combining a
steep number function dn/dN∗c with the uncertainties in
estimates of N∗c. Second, when optimizing the likelihood
over the width of the velocity filter σc, we included a prior
on the likelihood of the form ln(σc/σ0) with the arbitrary
normalization of σ0 = 103 km/s. This prior makes the like-
lihood independent of σc when a candidate cluster contains
only one galaxy with a redshift measurement3. Third, we
added a weak prior on the relation between N∗c and σc
based on our initial results. Preliminary, consistent mod-
els found that log N∗c = 1.13+1.90 log(σc/1000km/s) with
a dispersion of 0.39 dex. After determining N∗c as a func-
tion of σc on a coarse grid, the likelihoods were weighted
by this log-normal prior. The prior is only important for
setting the filter width σc in systems with too few galax-
ies for a direct estimate of the velocity dispersion. It also
helps to suppress the superposition of clusters into a single
larger cluster because superpositions have large apparent
velocity dispersions for the number of galaxies. Fourth, for
the trial galaxy used to compute the likelihood we used the
mean surface density inside the break radius rc rather than
the central density of the profile in computing the likeli-
hood. This reduces a bias towards finding clusters centered
on isolated galaxies given our cusped density profile.
3.5. Matching to Existing Catalogs
Once we have identified a cluster candidate we want to
compare its properties either to the known properties of
our synthetic catalog or to other properties of the real
clusters. We cross reference our cluster catalog to either
published or the synthetic cluster catalogs both through
the member galaxies and by matching the cluster coor-
dinates and redshifts, separately tracking both identifica-
tions. Each galaxy is associated with its most probable
cluster, the cluster which maximizes its membership like-
lihood, and has a likelihood ratio δi between the probabil-
ity that it is a member of that most probable cluster and
the probability that it is a field galaxy. This likelihood
ratio defines the probability that the galaxy is a cluster
member, pi = δi/(1 + δi).
In matching our new catalog to existing catalogs we re-
quire algorithms which are fully automated and robust.
They will not be perfect, but we cannot afford to conduct
laborious individual case studies in a broad survey. For
our synthetic data, each galaxy is labeled by its parent
cluster, allowing us to match the output and input cata-
logs by identifying the most common parent cluster for all
the galaxies with membership probability pi > 0.5 (δi > 1)
in each output catalog. This means of identification is ro-
bust, with relatively few multiple identifications or false
3For a maximum likelihood estimate of the velocity dispersion
from a set of velocity measurements vi, adding this prior leads to an
unbiased estimator of the velocity dispersion. If − ln L = N ln σ +
Pi(vi − ¯v)2/(2σ2) adding a ln σ prior and solving for σ leads to the
unbiased estimator σ2 = (N − 1)−1Pi(vi − ¯v)2. Without the prior
the estimator is biased by N/(N − 1).
5
identifications.
For the real data, the inhomogeneity of local cluster
identification enormously complicates cross references be-
tween our cluster identifications and previous studies. The
primary problem is that cross referencing is best done by
cross referencing the proposed member galaxies, but few
local group/cluster surveys supply such data. We settled
on an ecumenical approach using the NASA Extragalac-
tic Database (NED) to identify the galaxies likely to be
associated with each known cluster and then matching
our cluster catalog to the existing catalogs based on the
member galaxies. We searched NED for all "clusters" (ob-
ject types GClstr, GGroup, GTrpl & GPair). From this
list of 31130 clusters we looked for overlapping identifica-
tions which were not recognized by NED. This required
some iteration, and was a hopeless task at low redshift
(near Virgo) where a nearly infinite number of "groups"
have been found, many of which appear to be substruc-
tures of a larger object. Based on these identifications we
assigned cluster identifications to galaxies within a pro-
jected radius 0.5h−1 Mpc and a redshift difference less
than 1000 km/s. We then attempted to match our cluster
detections to these galaxies using both the match to the
cataloged cluster position and redshift and the most com-
mon cluster names associated with the galaxies we assign
to our detection. The final cross-matching system works
reasonably well but is by no means perfect.
3.6. Mass Estimates
In addition to identifying the parent cluster and the
member galaxies we would also like to estimate the mass
of the cluster. We have four different ways of making the
comparison. First, for the synthetic catalogs we have di-
rect estimates of the masses. Second, for the real data
we have measurements of X-ray luminosities and temper-
atures which may serve as surrogate estimates of the mass.
Third, we have the characteristic number of galaxies in the
clusters. Fourth, we have estimates of the galaxy velocity
dispersion in the clusters.
Because clusters lack sharp edges, we must exercise
some care in defining and comparing our mass estimates
(White 2001,2002). In fact, we found that failure to prop-
erly match mass definitions affects both the normalization
and the slope of our estimates of the halo multiplicity func-
tion. For this reason we give explicit details of our proce-
dure below.
We use two standard mass estimates for the clusters in
our synthetic catalog. The friends-of-friends mass, MF oF ,
is the total mass of the N-body particles linked into a
group using the FoF algorithm with a linking length of
b = 0.2 in units of the mean inter-particle spacing. We as-
sign galaxies to the clusters based on the FoF mass (Eq. 4)
because it can be computed quickly while we are build-
ing the synthetic galaxy catalogs. We also calculate sep-
arately the mass M200 inside the radius rM200 for which
the enclosed density is ∆M = 200 times the critical den-
sity. When we compare these two estimates we find that
M200 ≃ 0.66MF oF because the FoF mass corresponds to a
lower density threshold of ∆M ≃ ΩM /b3 ∼ 40.
For the real data, where we do not measure the masses
directly, we will compare our estimates from the galaxies
to X-ray luminosities and temperatures. We used compi-
lations of X-ray luminosities from Bohringer et al. (2000),
6
Cruddace et al. (2001), Ebeling et al. (1996, 1998, 2000),
De Grandi et al. (2000), Mahdavi et al. (2000), and
Reiprich & Bohringer (2002), and compilations of X-
ray temperatures from David et al. (1993), Ponman et
al. (1996), Markevitch (1998), Helsdon & Ponman (2000),
Ikebe et al. (2002) and White (2000). Most of the X-ray
luminosity data is drawn from the ROSAT All-Sky Sur-
vey (RASS, Voges et al. 1999). We rescaled the X-ray
luminosities and fluxes to a common scale by allowing for
logarithmic offsets relative to Reiprich & Bohringer (2002)
for each survey, with the offsets determined from clusters
with measurements in multiple surveys. The offsets we
applied are listed in Table 1. Measured scaling relations
for clusters of galaxies are (Markevitch 1998)
log(cid:18) LXh2
1044ergs/s(cid:19) = (0.15±0.04)+(2.10±0.24) log(cid:18) T
6keV(cid:19)(18)
and (Horner, Mushotsky & Scharf 1999; Nevalainen, Marke-
vitch & Forman 2000; Finoguenov, Reiprich & Bohringer 2001;
Xu, Jin & Wu 2001)
(19)
M∆ ≃ 1015h−1M⊙ (cid:18) TX
1.3keV(cid:19)3/2(cid:0)∆M E2(cid:1)−1/2
where ∆M is the density threshold used to estimate the
mass M∆, and E(z) ≡ H(z)/H0 = 1 for our low redshift
sample. The normalization of this last relation is currently
quite uncertain, at the 30-50% level, due to the notori-
ous difficulties inherent in measuring the mass of a cluster
(e.g. Evrard, Metzler & Navarro 1996) and differences in
methods for fitting temperatures to the X-ray emitting
ICM. (See Fig. 2 of Huterer & White 2002 for a compi-
lation of theoretical and observational results.) Different
authors use different methods to model the 'temperature'
of the plasma, different methods of estimating the mass
and indeed even different definitions of the mass! To cloud
the picture even further, numerical simulations based on
purely adiabatic gas physics find a different normalization
of this relation than most of the observations (although
it must be noted that significant progress in this regard
has occurred recently, e.g. Muanwong et al. 2002, Voit et
al. 2002).
The number of member galaxies or their aggregate lu-
minosity must be carefully defined before we can compare
it to other results or theoretical models. Our fit parameter
N∗c is defined to be the number of L ≥ L∗ galaxies inside
the fixed, spherical outer radius rout = crc = 0.8h−1 Mpc.
Postman et al. (1996, 2002) and Donahue et al. (2001) use
the luminosity, measured in units of L∗, rather than the
number. We prefer the number of galaxies, N∗, to this
equivalent luminosity, Λ∗, because it has a closer relation
to the Poisson statistics which determine the detectability
of clusters and may depend less on the defining luminosity
function or wavelength. For the same normalizing aper-
ture, the two quantities are related by
Λ∗L∗ = L∗N∗Γ[2 + α]/Γ[1 + α, 1] = 5.0N∗L∗
(20)
for α = −1.09. We cannot precisely match the Postman et
al. (1996) definition of Λcl because we use different filters
and luminosity functions. To the extent that Λcl is defined
by a rcl = 1.1h−1 Mpc (1.5h−1
75 Mpc) radius aperture, we
expect N∗cl ≃ (g(rcl/rc)/F (c))N∗c = 1.6N∗c.
While N∗c and Λcl are natural variables for finding clus-
ters, they are poor choices for comparing to theoretical
models. For comparisons to theoretical models we will use
the number of L > L∗ galaxies N∗∆ inside the spherical
radius rN ∆ such that the enclosed galaxy density is ∆N
times the average density of galaxies. Given a cluster with
N∗c galaxies brighter than L∗ inside rout we determine r∆
and N∗∆ by solving the defining relations
N∗∆ ≡ N∗cF (rN ∆/rc)/F (c) =
4π
3
n∗∆r3
N ∆Γ[1 + α, 1].
(21)
For reasons of computational efficiency and stability we
keep c fixed when we determine rN ∆. In most theoretical
models, the distribution of galaxies basically traces the
distribution of mass (Katz, Hernquist & Weinberg 1999;
Gardner et al. 1999; Pearce et al. 1999; White, Hernquist
& Springel 2001). However, if we assume the galaxies are
unbiased4, to isolate the same physical region we must set
∆N = ∆M /ΩM because the galaxy overdensity is defined
relative to the average galaxy density while the mass over-
density is defined relative to the critical density rather
than the mean mass density. Thus, taking ΩM = 0.3,
the number of galaxies brighter than L∗ inside the re-
gion with the standard mass overdensity of ∆M = 200
is N∗666 = 0.66N∗F oF with rM200 = rN 666.
Our best theoretical estimate of the conversion to Post-
man et al. (1996, 2002) and Donahue et al. (2001) is
Λcl ≃ 11N 3/4
∗666
(22)
for Λcl >∼ 10. The slope difference is created by the dif-
ference between a fixed rcl = 1.5h−1
75 Mpc aperture for
Λcl and a variable aperture rN 666 for N∗666.
In prac-
tice, both the normalization and slope could differ from
this estimate because of differences in the luminosity func-
tions, wavelengths (K band versus I band) and matched
filter structures. A simple check is to compare Λcl and
N∗666 values for clusters of different Abell (1958) cluster
richness classes. Our overlap with the Abell (1958) cat-
alog is limited because the average redshift of the Abell
clusters is significantly higher than our 2MASS catalog.
A quick match based on NED cluster names found 31,
39, and 15 Abell richness class 0, 1 and 2 clusters with
fractional errors in N∗666 smaller than 25% in the cata-
logs of §5. We find an average richness and dispersion of
hN∗666i = 5.3± 2.7, 8.9± 4.3 and 13.4± 6.4 for Abell rich-
ness classes of 0, 1 and 2 respectively. The ratios of the av-
erage N∗666 values roughly track the ratios of the number
of galaxies associated with each richness class. Postman
et al. (2002) estimate that Λcl = 50, 80 and 130 should
correspond to the typical Abell richness classes 0, 1 and
2 clusters which differs significantly from our estimated
conversion to Λcl (Eq. 22) of Λcl = 41 ± 16, 60 ± 22, and
84 ± 28 for the same richness classes. While using Abell
richness classes to compare modern algorithms can hardly
be recommended, the comparison suggests that we should
explore empirical conversions between Λcl and N∗666.
We find clusters with richnesses as low as N∗666 ≃ 0.1
in our catalogs, and this requires some explanation. First,
4The 'galaxies' in the simulated catalog are very nearly unbiased.
The bias of the 2MASS galaxies is currently unknown, but we shall
argue later it is likely they are relatively unbiased.
7
Adopted Offsets Between X-ray Data
Table 1
Name
Offset
Error Ncluster
BCS
de Grandi
Cruddace
Ponman
RASSCALS
HIFLUGCS
XBACS
NORAS
0.036
0.246
0.039
−0.084
0.067
≡ 0
0.040
0.051
0.005
0.009
0.008
0.035
0.008
≡ 0
0.004
0.004
351
145
109
9
55
193
377
299
Note. -- The offsets, defined by the loga-
rithmic correction needed to match the fluxes
quoted by Reiprich & Bohringer (2002), ap-
plied to the X-ray luminosities from the dif-
ferent samples we analyzed. The offsets were
generally small, with only that applied to the
sample of de Grandi et al. (2000) being signif-
icant.
Estimates of Halo Occupancy Function
Table 2
Data
Mass Scale
Correlation
Transformations
zero-point A
slope B
Comments
Synthetic True
M200
σ200(M200), Eq. 28 N∗666(σ)
dn/dM200
dn/dN∗666
N∗666(M200)
2MASS
σ200(M200), Eq. 28 N∗666(σ)
N∗666(TX )
σ(TX ), Eq. 32
TX (M200), Eq. 19
dn/dM200
N∗666(LX )
Average
N∗666(σ)
N∗666(TX )
N∗666(LX )
Average
dn/dN∗666
σ(LX ), Eq. 31
σ(TX ), Eq. 32
LX (TX ), Eqs. 18
LX (TX ), Eqs. 34
Adopted Standard
1.32
1.25 ± 0.03
1.23 ± 0.04
1.43 ± 0.05
1.37 ± 0.05
1.43 ± 0.09
1.28 ± 0.03
1.24 ± 0.05
1.25 ± 0.05
1.22 ± 0.04
1.25 ± 0.03
1.57 ± 0.04
1.60 ± 0.07
1.59 ± 0.06
1.56 ± 0.06
1.56 ± 0.06
1.58 ± 0.02
1.49 ± 0.04
1.49 ± 0.04
1.54 ± 0.05
1.44 ± 0.17
0.98
Bright X
Bright X
0.89 ± 0.02
0.94 ± 0.03
1.13 ± 0.04 Nthresh = 3
1.04 ± 0.04 Nthresh = 5
1.09 ± 0.08 Nthresh = 10
0.90 ± 0.02
1.13 ± 0.09
1.06 ± 0.06 All X
0.98 ± 0.06
1.02 ± 0.10
1.10 ± 0.02
1.39 ± 0.11
1.31 ± 0.08 All X
1.05 ± 0.07
1.19 ± 0.08
1.21 ± 0.14
1.13 ± 0.03 Nthresh = 3
1.08 ± 0.03 Nthresh = 5
1.13 ± 0.06 Nthresh = 10
1.10 ± 0.09
Bright X
Bright X
Bright X
Note. -- The halo occupancy function has the form log N∗666 = A + B log(M200h/1015M⊙). Except for the fits to the
cluster number function, all results are for a standard cluster sample consisting of the systems with at least Nv ≥ 5
associated redshifts. The cluster number function estimates are for threshold galaxy numbers of Nthresh = 3, 5 and 10.
The halo mass function dn/dM200 is that of Jenkins et al. (2001) converted to M200 from the original FoF masses assuming
M200 = 0.66MF oF as found in our simulations for the same FoF linking length. The N-body simulations used to generate
our synthetic catalog have the same initial parameters. Models based on X-ray data samples limited by TX ≥ 1 keV and
LX ≥ 1042h2 ergs/s contain the comment "Bright X". The Average entries are simply the average coefficients and their
scatter. The Adopted standard combines the 2 Average relations and the Nthresh = 5 result for the 2MASS data.
8
low values of N∗666 only imply that it is unlikely the clus-
ter contains any bright (L > L∗) galaxies.
It does not
imply the cluster is unlikely to contain galaxies. For ex-
ample, an N∗666 = 0.3 cluster should contain 3 (7) galaxies
with L > L∗/10 (L > L∗/100) inside its virial radius. We
can routinely detect such systems provided the redshift is
low enough for us to detect the lower luminosity galaxies.
Second, we detect the systems in projection, and the num-
ber of galaxies projected inside the virial radius is larger
than the number inside the spherical virial radius. This is
a modest effect for the NFW profile, but real clusters are
embedded in more extended infall regions which boost pro-
jection effects. Third, there are significant errors in these
low estimates of N∗666 because there are so few galaxies.
Thus, these low richness clusters can be real virialized sys-
tems, but they must be treated with care.
Finally, we have estimates of the velocity dispersion of
the clusters. Although we optimized the cluster velocity
width σck as part of our matched filter, we do so at low
resolution (150 km/s) and it includes contributions from
galaxies at large distances from the cluster center. To bet-
ter mimic a virial velocity for a cluster we use a member-
ship probability weighted estimate of the velocity disper-
sion. We take all galaxies with measured redshifts, mem-
bership probabilities pi ≥ 0.5 and projected separations
less than the model outer radius R ≤ crc = 0.8h−1 Mpc
and estimate a mean redshift
dn
dN
=
dn
dM (cid:12)(cid:12)(cid:12)(cid:12)
dM
dN(cid:12)(cid:12)(cid:12)(cid:12) .
If we assume that the cluster mass function is known, then
we can infer the halo occupancy function by determining
the function N (M ) needed to transform the assumed mass
function into the observed number function. We use the
Jenkins et al. (2001) CDM mass function, which should
closely match that of our N-body simulation since it is
based on numerical simulations with similar cosmologi-
cal parameters (see White 2002 for further discussion).
and a velocity dispersion
hczi =X czipi/X pi
(P pi)2
(P pi)2 −P p2
P pi
i P(czi − hczi)2pi
σ2
0 =
which we then correct to the rest frame σ = σ0/(1 + z).
The leading term plays the role of the familiar N/(N − 1)
factor needed to make the variance of the velocities about
the mean an unbiased estimator of the velocity dispersion.
The weighting by membership probability plays an equiva-
lent role to the clipping of outliers in standard approaches
to estimating cluster velocity dispersions (see discussion
in Borgani et al. 1999). and the method is similar to that
used by Barmby & Huchra (1998).
3.7. The Cluster Number Function
We can also estimate the cluster number function, dn/dN ,
which is the distribution of clusters with respect to the
number of member galaxies.
If the number of member
galaxies, N (M ), is a simple function of the cluster mass
then the number function is simply related to the cluster
mass function dn/dM by
(23)
(24)
(25)
1
0.8
0.6
0.4
0.2
0
10
100
1000
Fig. 1. -- Completeness as a function of ∆ ln L(0) for clusters
with Mcut ≥ 1014h−1M⊙ and cz ≤ 10000 km/s. The solid line
shows the completeness, the dashed line shows the fraction of clus-
ter candidates which are false positives. The upper dotted line (at
∆ ln L(0) = 100) shows the fraction of cluster candidates which are
real clusters with masses Mcut/3 ≤ M < Mcut. The lower dotted
line (at ∆ ln L(0) = 100) shows the fraction of cluster candidates
which cannot be identified with any halo.
1
0.8
0.6
0.4
0.2
0
0
5000
Fig. 2. -- Completeness as a function of redshift for clusters
with Mcut ≥ 1014h−1M⊙ and ∆ ln L(0) = 150. The solid his-
togram shows the fraction of the M ≥ Mcut clusters detected in
each bin, the dashed histogram the fraction of candidates which
are false positives. The dotted histogram, decreasing with redshift,
shows the fraction of cluster candidates which are real clusters with
Mcut/3 ≤ M < Mcut, while the rising dotted histogram shows the
fraction of candidates which cannot be identified with any halo.
The vertical lines mark the redshifts where the average M = Mcut
cluster has 10 or 3 galaxies.
The Jenkins et al. (2001) mass function was derived us-
ing MF oF masses for the clusters, so we transformed it
from dn/dMF oF to dn/dM200 assuming the average rela-
tion that M200 = 0.66MF oF we observe in our simulations
for the same linking length. This method is similar to the
approach used by Berlind & Weinberg (2002) and Berlind
et al. (2002).
We can estimate the number function using the V /Vmax
method (e.g. Moore et al. 1993) as follows. At redshift z,
a cluster normalized by N∗666 contains
Ng(z) = N∗666Γ[1 + α, Llim(z)/L∗]/Γ[1 + α, 1]
(26)
galaxies, where Llim(z) is the luminosity corresponding to
the survey magnitude limit at redshift z. Suppose we de-
tect all clusters with at least Ng ≥ Nthresh galaxies. Then
we detect clusters in the volume Vmax(N∗666) defined by
the comoving survey volume out to the redshift zlim which
solves Nthresh = Ng(zlim). The integral number function
is then
(27)
n(> N ) = XN ≥N
Vmax(N )−1
where the sum includes only clusters with more than the
threshold galaxy number at their observed redshift. For
large values of Nthresh we are less sensitive to the Poisson
errors in estimating N∗666 and minimize the risk of includ-
ing false positives as clusters. This comes at the price of
increased Poisson errors in the number function, because
we include fewer clusters, and increasing sample variance
because we include less volume. We limited the sample
used and Vmax to cz < 15000 km/s and used bootstrap
resampling of the catalog to estimate the uncertainties in
the number function.
4. TESTS WITH THE SYNTHETIC CATALOG
We start by testing the algorithm and our recovery of
the halo occupancy function in the synthetic data designed
to mimic the real 2MASS galaxy sample.
4.1. Completeness and contamination
We first consider the completeness of the resulting cata-
logs. As we discuss in WK, the likelihood associated with a
cluster roughly scales with the number of member galax-
ies, so the distance at which we can detect clusters can
be roughly estimated from the number of galaxies.
In
our models, the average number of cluster galaxies drops
to NF oF = 10 (3) at redshifts of cz = 12000 (17000),
7400 (13000), and 2700 (7500) km/s for clusters of mass
M h/M⊙ = 3 × 1014, 1014 and 3 × 1013 respectively. The
spread in the likelihoods for a fixed mass and redshift will
be considerable because of the Poisson variations in the
number of galaxies bright enough to be visible and be-
cause of the random variations in the numbers of measured
redshifts for the fainter galaxies. In WK we showed that
the cluster likelihood essentially scales with the number
of member galaxies, so that by scaling the selection likeli-
hood as ∆ lnL(z) = C(z)∆ lnL(0) we can extract clusters
with a nearly constant mass threshold as a function of red-
shift. The choice of the zero-redshift likelihood threshold,
∆ lnL(0), is related to the mass cut and its completeness.
Fig. 1 illustrates this for model clusters with M ≥
1014h−1M⊙. Of the 101 clusters with M ≥ 1014h−1M⊙
9
and cz < 104 km/s, we miss only 5 at our likelihood cut-
off. Of these, one was merged into another cluster, 4
were systems with M ≃ 1014h−1M⊙ and fewer galaxies
than expected given their mass. We can achieve very high
completeness out to the redshift where the typical cluster
contains only 3 galaxies. However, all samples with high
completeness also have high false positive fractions. If the
likelihood threshold is kept high enough, the vast majority
of these false positives are real clusters with masses just
below the cutoff mass scale -- very few of these false pos-
itives lack identification with any M > 3 × 1013h−1M⊙
halo. For lower likelihood thresholds the mass range of
the false positives broadens and then slowly becomes dom-
inated by unmatched systems where we could not identify
the detection with a cluster in the input catalog.
Fig. 2 shows the redshift dependence of the complete-
ness and false positive rates after choosing ∆ lnL(0) to
have a 50% false positive rate. The completeness declines
slowly with redshift, but remains close to 90% complete
from 0 ≤ cz ≤ 10000 km/s. The false positive rate aver-
ages 50%, and the false positives consist almost entirely
of real clusters with masses between 3 × 1013h−1M⊙ and
1014h−1M⊙. False positives that correspond to no clus-
ter become important only for redshifts beyond which a
1014h−1M⊙ cluster contains fewer than 3 galaxies. Within
our cz ≤ 10000 km/s selection window there is a total vol-
ume of 4.1V0 where V0 = (100/h)3 Mpc3, containing 101
clusters more massive than 1014h−1M⊙. Using these clus-
ters to estimate the mass function would be limited by the
correction for the contamination by lower mass clusters
rather than the correction for completeness or statistical
uncertainties including sample variance.
The true completeness of the recovered cluster sample
10
1
0.1
Fig. 3. -- The number of galaxies N∗666 as a function of the virial
mass M200. The points show all detected clusters with at least
Nv ≥ 5 associated redshifts which have been matched to the input
cluster catalog. The heavy solid line shows the true relationship and
the light dashed lines show the estimated relation and its width as
estimated from the variance of the points.
10
is higher than we have quoted because of the way we have
generated the synthetic catalog. In regions of high den-
sity, where we expect the most massive clusters, the FoF
algorithm with a linking length of 0.2 is prone to merg-
ing sub-clumps into larger objects. Our search algorithm
instead detects the individual sub-clumps with lower like-
lihoods than would be expected for an object as massive
as the FoF cluster. The sub-clumps may have too low a
likelihood to be included (creating a false impression of in-
completeness) or multiple clumps may be included (adding
to the false positives). Most peculiar high FoF mass but
low detection likelihood systems could be traced to this
feature of the FoF algorithm.
4.2. Halo occupancy function
Given a well-defined cluster catalog, the next step is to
examine how well the cluster parameters agree with the
true properties of the clusters. We first test our ability to
recover the true halo occupancy function N (M ) given our
knowledge of the halo masses for the synthetic catalogs.
For the real 2MASS data we will consider cluster X-ray lu-
minosities and temperatures or galaxy velocity dispersions
as surrogate, but independent, estimates for the mass.
We test our ability to recover the halo occupancy func-
tion by comparing the number of galaxies N∗666 inside the
∆N = 666 contrast level. This corresponds to the num-
ber of galaxies inside a mass contrast level of ∆M = 200
and should, on average, match 0.66N∗F oF . There are two
critical issues to estimates of the halo occupancy function.
First, definitions are critical! The "number of galax-
ies in a cluster" must be precisely and similarly defined
both for the algorithm used to find the clusters and the
theoretical model used to interpret the results. Failure to
match the definitions, for example by using N∗c instead of
N∗666, makes it impossible test the algorithm (for our syn-
thetic catalog) or to interpret the results (for the 2MASS
catalog).
In particular, N∗c > N∗666 for low mass clus-
ters where rM200 < crc = 0.8h−1 Mpc and N∗c < N∗666
for high mass clusters where rM200 > crc = 0.8h−1 Mpc.
Thus, the slope of N∗c(M ) ∼ M 0.75 is significantly flatter
than the slope of N∗666 ∼ M . We expect the same flatten-
ing of the slope will be found for Λcl as used by Postman
et al. (1996, 2002) and Donahue et al. (2001) for similar
reasons.
Second, in estimating relations like N (M ) it is impor-
tant to select subsamples which minimize the introduction
of mass and number-dependent biases. For example, esti-
mating N (M ) using a sub-sample selected such that the
standard error in N is less than a threshold or that the
number of galaxies in the cluster exceeds some threshold
will preferentially include low mass clusters with high val-
ues of N over those with low values of N . This biases
estimates of N (M ) to have flatter slopes than the true
relation.
With the logarithmic prior for N∗c included in the like-
lihood, our ability to estimate the halo occupancy be-
comes very robust. Simple redshift cuts work well, but
we will define our standard sample by systems which con-
tained at least Nv ≥ 5 galaxies with redshifts. After
selecting a cluster sample we fit the distribution as a
power law including the formal uncertainties in the es-
timate of N∗666 for each cluster. The results, illustrated
in Fig. 3, are remarkably good. Above 1013h−1M⊙ the
input relation is well approximated by a power law (see
log N∗666 = A + B log(M200h/1015M⊙) with
Table 2),
1000
100
10
1
0.1
100
1000
100
1000
Fig. 4. -- Comparison of the velocity dispersion σ estimated for
the cluster candidate to the velocity dispersion σ200 of the dark
matter particles inside r200. Cluster candidates including at least
Nv ≥ 10 (5 ≤ Nv < 10) redshifts are shown as the solid (open)
points. The solid line indicates σ = σ200 while the dashed line is
the best-fit power law. The overall correlation is excellent, but there
are outliers due to contamination.
Fig. 5. -- The number of galaxies N∗666 as a function of the
estimated cluster velocity dispersion σ for the synthetic data. Filled
(open) points have Nv ≥ 10 or (5 ≤ Nv < 10) velocities contributing
to the dispersion estimate. The heavy solid line shows the true
scaling, and the dashed lines show the best fit power law and its
width as estimated from the variance of the points.
A = 1.32 and B = 0.98. With our standard sample we
find A = 1.25 ± 0.03 and B = 0.89 ± 0.02. The error bars
are small because of the large number of clusters (484), so
the fits formally disagree with the input relation by 3 -- 4
standard deviations.
The uncertainties in the halo occupancy function are
dominated by systematic errors arising from sample se-
lection rather than statistical errors. For example, if we
simply include all systems with cz ≤ 15000 km/s, roughly
the limit for detecting L∗ galaxies, we find parameters
(A = 1.20 ± 0.02 and B = 0.92 ± 0.02) which has a lower
normalization, a steeper slope and more scatter about the
mean relation. For systems with cz ≤ 10000 km/s, we find
A = 1.23±0.02 and B = 0.95±0.01. If we use a likelihood-
dependent redshift cut such as cz < 10000 log(lnL) km/s,
we find A = 1.26 ± 0.02 and B = 0.95 ± 0.01. We will
adopt the Nv ≥ 5 sample as our standard from this point,
and summarize the estimates of the occupancy function in
Table 2.
In summary, we can recover the true halo occupancy
function of the synthetic catalog reasonably well. The un-
certainties are dominated by systematic errors at the level
of 0.05 in the slope and 10% in the normalization of the
relation. Some of these problems are due to the generation
of the synthetic catalog rather than the search for clusters
in the synthetic catalog. The FoF algorithm used to find
clusters in the N-body simulation is prone to linking ob-
jects which both the eye and our matched filter will regard
as neighboring but separate objects. These then appear in
our test as objects with high 'true' mass but low 'recov-
ered' mass, possibly creating a modest bias towards flatter
slopes and lower normalizations. While this aspect could
be improved upon, we will leave exploration of 'unlinking'
algorithms to future work.
Unfortunately in the real data we do not know the in-
trinsic masses of the halos and must use a surrogate esti-
mate such as the cluster velocity dispersion, X-ray lumi-
nosity or X-ray temperature. The surrogate we estimate
as part of our algorithm is the cluster velocity dispersion
σ, and in Fig. 4 we compare our estimated velocity disper-
sions σ based on Eq. (24) to the velocity dispersion of the
dark matter particles σ200,DM inside rM200 of the cluster
center. We select the sample for comparison with both
our standard limits on the redshift (cz < 15000 km/s)
and a minimum number Nv ≥ 5 of velocities to be in-
cluded in the redshift estimate. Fit as a power law with
(σ/103km/s) = a(σ200,DM /103km/s)b we find little offset,
a = 0.93 ± 0.03, or slope difference, b = 1.04 ± 0.04. Re-
stricting the analysis to systems with Nv ≥ 10 produces
similar results, with a = 0.89 ± 0.05 and b = 0.94 ± 0.06.
The scatter is consistent with the uncertainties in the ve-
locity dispersion estimates.
We can also estimate N∗(σ), as shown in Fig. 5. In the
simulations the velocity dispersion is related to the mass
by
σ200 ≃ 925(cid:18) M200h
1015M⊙(cid:19)0.34
km/s
(28)
with a 10% dispersion about the power-law. The power
law describing the input N∗666(σ200) relation has N0 =
25.6 and x = 2.75 for N∗666 = N0(σ/103km/s)x. The ob-
served relation, with our standard selection procedures, is
nearly identical, with x = 2.76± 0.08 and N0 = 21.1± 2.0,
11
to the input relation. The uncertainties are large because
both quantities have relatively large statistical uncertain-
ties. These parameters change little as we adjust the sam-
ple selection criteria. Alternatively, we can use Eq. (28) to
transform from σ to M200 to estimate the occupation num-
ber directly, with results very close to the input relation
(see Table 2).
4.3. The Cluster Number Function
Finally we examine how well we can recover the clus-
ter number function dn/dN∗666 of the synthetic catalog.
Fig. 6 shows the distributions estimated using thresholds
of Nthresh = 3, 5 and 10 in defining Vmax (see §3.7). The
three estimates are mutually consistent, extending over
two decades in N∗666. We can also compute the distri-
bution for the (200h−1Mpc)3 N-body simulation volume
from which our synthetic catalog is drawn. The charac-
teristic volume for fair samples of the universe is roughly
V0 = (100h−1Mpc)3, which we shall use as a unit of
volume.
In these units the simulation has a volume of
8V0, and to generate the synthetic catalogs we periodi-
cally replicate the simulation. The number function of
low N∗666 clusters is derived from a small region of the
simulation, while the number function of high N∗666 clus-
ters is drawn from several of the periodic repetitions of
the data. There is also good agreement with the semi-
analytic number function found by combining the the syn-
thetic catalog's multiplicity function with the Jenkins et
al. (2001) CDM mass function following the procedures in
§3.7 (Eq. 25). The only significant difference is the absence
of high N∗666 clusters, but this is a real feature of the N-
0.1
1
10
100
Fig. 6. -- The number function dn/dN∗666 for the synthetic data.
The triangles, squares and pentagons show the results for Nthresh =
3, 5 and 10 and their statistical errors. The solid histogram shows
the actual distribution in the (200h−1Mpc)3 N-body cube and its
Poisson errors for the same bins. The smooth solid curve shows
the expected distribution based on the Jenkins et al. (2001) CDM
mass function and the synthetic catalog's multiplicity function. The
mass scale at the top is based on the true occupation function for
the synthetic data.
12
body simulation used to produce the synthetic catalog. If
we combine the mass function and the number function
to estimate the occupancy function, we obtain mutually
consistent estimates of N∗666(M200) for Nthresh = 3, 5 and
10 which have slopes and normalizations slightly above
(1 -- 2σ) the true relation (see Table 2).
5. 2MASS CLUSTERS
With this understanding of the algorithm, we can now
search for clusters in the real 2MASS data. We present no
catalog at present pending a reanalysis of the full 2MASS
data to a fainter limiting magnitude than is presently avail-
able. Instead, we focus on the properties of the clusters.
In any fitted relation, we have selected the cluster sam-
ple used to perform the fit in exactly the same manner
as was used for the synthetic catalog, and the order of
the analyses parallels that used to test the algorithm on
the synthetic catalog. Our standard sample consists of all
galaxies identified from the 2MASS extended source cata-
log (see Jarrett et al. 2000) with b > 5◦ to an extinction
corrected magnitude limit of K≤ 12.25 mag.
5.1. Tests of Completeness
We start with two tests of the completeness of the algo-
rithm for the real data. The first test compares the "shal-
low" cluster catalogs generated using a bright subsample
of the galaxy catalog, truncated at K≤ 11.25 mag, to the
"deep" cluster catalogs generated using the complete sam-
ple to K≤ 12.25 mag. The second test is to compare our
cluster catalog to local X-ray surveys.
0
Fig. 7. -- Detection of X-ray clusters as a function of redshift
and X-ray flux fX . The solid points are X-ray clusters that were
found and the open points are X-ray clusters which were missed.
The squares show the clusters from the HIFLUGCS survey which is
complete to fX = 2 × 10−11 ergs/cm2 s (the horizontal line). The
triangles show all clusters with X-ray flux measurements that we
matched to our output catalog. The missed HIFLUGCS target at
low redshift is the NGC 4636 group, and it is probably a failure in
our matching methods rather than a true non-detection.
5.1.1. Comparison of shallow and deep surveys
We first discuss the results from comparing the shallow
and deep cluster catalogs. With a one magnitude change in
the survey depth, individual clusters will have significant
changes in their membership, as a cluster which contained
only L > L∗ (L > 0.1L∗) galaxies in the bright subsample,
would on average contain 3.4 (1.5) times as many galax-
ies in the complete sample. Using the same termination
criterion for the search, the shallow catalog contains only
375 clusters, as compared to 1743 in the deep catalog.
We first consider the reliability of the catalog by trying
to identify shallow catalog clusters which have no coun-
terpart in the deeper catalog. By cross-referencing the
cluster assignments of the galaxies in the two catalogs we
estimated a matching probability for identifying a cluster
in the shallow catalog with each cluster in the deep cata-
log. We included all galaxies with a 50% or higher cluster
membership probability. The cluster matching probability
can range from 0, if no galaxy assigned to a shallow cluster
has been assigned to a deep cluster, to unity if every galaxy
assigned to the shallow cluster is assigned to the same deep
cluster. Only 4 of the 375 clusters lack a firm counterpart
in the deeper catalog (defined by a matching probability
< 50%), and an additional 8 may have ambiguous assign-
ments (matching probability of 50 -- 70%). Most shallow
clusters have unambiguous counterparts, with 78%, 85%
and 91% of the shallow clusters having match probabili-
ties exceeding 95%, 90% and 85% respectively. The match
probabilities are not unity because galaxies with modest
membership probabilities can be dropped or assigned to
a different cluster based on the deeper catalogs, reducing
the match probability, even if the galaxies in the cluster
cores remain perfectly matched. Thus, if we define a false
positive as a cluster which will lack a clear counter part
in a deeper survey, our method appears to have a false
positive rate of 1 -- 3% depending on the desired threshold.
Be warned, however, that this false positive rate is not
the same as the true false positive rates we considered for
the synthetic data -- a real but unvirialized overdensity of
galaxies can be identified as a false positive in the syn-
thetic data, but our deeper survey may still identify the
clump as a cluster.
Next we examine the relative completeness of the two
catalogs. We divided the deep catalog into bins of N∗666
and computed the fraction of the deep catalog clusters
found by the shallow survey as a function of redshift.
For each richness bin, the shallow survey is at least 80%
complete until the redshift at which the least rich clus-
ters in the bin are expected to have fewer than 3 galax-
ies. For example, for the clusters in the deep catalog with
1 < N∗666 < 3, corresponding to masses near 1014h−1M⊙,
the completeness of the shallow catalog is 92% (61%) for
redshifts below that where an N∗666 = 1 (3) cluster is ex-
pected to contain 3 galaxies. The deep catalog presumably
has similar properties.
The final comparison we make is to compare the esti-
mates of N∗666 for the 363 well-matched clusters from the
two catalogs. We find that
log N shallow
∗666 = (0.07± 0.01) + (1.00± 0.01) log N deep
∗666 (29)
with a variance of about the relation of 0.21 dex. Formally,
the error estimates for N∗666 would have to be raised by
14% to make the variance consistent with the error esti-
mates. The shallow catalog estimates of N∗666 are higher
by 17% on average (also found as the median difference).
The offset is probably a real systematic effect -- the shal-
low catalog will preferentially include clusters with excess
numbers of K < 11.25 mag galaxies because an upward
Poisson fluctuation increases the likelihood of including
the cluster in the catalog. The same cluster is unlikely to
show the same upward fluctuation in the K < 12.25 mag
catalog, producing a net bias. This interpretation pre-
dicts that a comparison restricted to clusters in the shallow
survey with more galaxies, corresponding to clusters with
smaller uncertainties in N∗666, should have less of a bias.
For example, if we restrict the comparison to the 41 clus-
ters in the shallow catalog with errors in log N∗666 smaller
than 0.10, the average offset is reduced to 7% (0.03± 0.02)
from 17%.
5.1.2. Comparison with X-ray catalogs
The second test compares our complete catalog to X-ray
selected samples. It is difficult to do so in a completely
satisfactory manner because the selection methods are so
different. We first compare our results to the 63 clus-
ters in the HIFLUGCS X-ray flux limited sample (fX >
2 × 10−11 ergs/cm2 s, Reiprich & Bohringer 2002). The
HIFLUGCS survey covers most of the sky with b > 20◦
plus small holes near the LMC, the SMC and the Virgo
cluster. Fig. 7 shows that we find all HIFLUGCS clusters
out to the redshifts where we expect even rich clusters
to have too few galaxies to be detected by our algorithm
(cz ≃ 15000 km/s). Our automated matching system iden-
tified counterparts to 97% (35 of 36) of the HIFLUGCS
clusters with cz < 15000 km/s. We missed one very low
redshift system, the NGC 4636 group at cz = 1320 km/s,
but closer examination strongly suggests that this is a
matching and coordinate problem rather than a genuine
failure. Between 15000 km/s and 20000 km/s we find
67% (8 of 12) of the HIFLUGCS clusters. We miss the
lower temperature clusters (hTXi = 4.3 keV) and find the
higher temperature clusters (hTXi = 6.6 keV). At higher
redshifts, an increasing fraction of X-ray clusters are gen-
uinely missed. Spot checks of these regions show only 0 -- 2
galaxies at the position and redshift of the X-ray cluster.
We can also use our N∗666 estimates to estimate an X-
ray flux (based on the correlation between N∗666 and the
X-ray luminosity in Eq. 36). This leads to 18 (14 below
cz = 15000 km/s) clusters whose N∗666 values and co-
ordinates suggest the cluster should have an X-ray flux
large enough to be included in HIFLUGCS. These flux
estimates are not very accurate, but we show their distri-
bution, in Fig. 7. They can be divided into three loose cat-
egories. First, there are 7 very low redshift systems with
cz < 2000 km/s. Some have very high likelihoods in our
catalog (the two most prominent are the Eridanus group
and the Ursa Major cluster which are the 26th and 38th
clusters found), but no published X-ray studies. These sys-
tems have virial sizes significantly larger than the ROSAT
field of view. Second, there are 5 at intermediate red-
shifts cz ∼ 5000 km/s. The two most prominent sys-
tems here, Abell S0805 and Abell 3574, are the 9th and
17th clusters found. Finally, there are 6 systems above
cz > 10000 km/s of which the most prominent are Abell
1913 and Abell S0740. After considerable and time con-
13
suming effort, we gave up on tracking down the origins
of these objects. Many of the systems are well studied
clusters with X-ray emission (e.g. Abell S0805) that do
not appear in any of the ROSAT catalogs we have used
for our comparisons. Some are systems contaminated by
X-ray emission from AGN (e.g. Abell 3574). The very
low redshift systems have angular extents comparable (or
larger) than the ROSAT aperture. In short, understand-
ing these systems in detail requires a careful analysis of
the ROSAT data which is beyond the scope of our present
effort. We note, however, that Donahue et al. (2002) also
find examples of optically detected clusters which appear
to have fainter than expected X-ray counterparts.
Fig. 7 also shows all other clusters for which we have
found a ROSAT X-ray flux measurement but are not de-
rived from complete, flux-limited samples. As these are
largely derived from the RASS survey as well, they tend
to have X-ray fluxes fX >∼ 2 × 10−12 ergs/cm2 s. The
automatic matches find 84% (178 of 213) of the clusters
with X-ray flux measurements in our input local cluster
catalog and cz < 15000 km/s. As with the HIFLUGCS
sample, some of the mismatches at low redshift are due to
cross-matching and coordinate problems.
In summary, by comparing our standard cluster catalog
to either a shallow catalog or X-ray cluster catalogs, we
can show that we achieve very high completeness up to the
redshifts where clusters contain too few galaxies for robust
detection (about 3 galaxies). It is difficult, however, to use
either comparison to robustly estimate a false positive rate.
The comparison of the shallow and deep catalogs can only
set a lower bound (of a few percent) on the false positive
rate. The HIFLUGCS sample sets an upper bound of 40%,
but this mainly reflects the limitations of the RASS and
our ability to estimate X-ray fluxes from N∗666 rather than
a true estimate of the false positive rate.
5.2. Further Checks of the Velocity Dispersion
Before proceeding to estimates of the halo occupancy
function, we made three additional checks of our velocity
dispersion estimates, comparing directly with other pub-
lished values and through comparisons to X-ray observa-
tions of the clusters. All fits of variable correlations are
performed as χ2 fits including the errors in both variables
and including all clusters with at least Nv ≥ 5 associated
redshifts that we could match to the other data.
First we checked our velocity dispersions against the ve-
locity dispersions σpub from Girardi et al. (1998; Table 2)
and Wu et al. (1999). We matched 98 clusters with at
least Nv ≥ 5 redshift measurements to these previous es-
timates, and we found good agreement within the rather
large errors (Fig. 8). If we fit a power-law, our σ is related
to σpub by
σ
1000km/s
= (1.09 ± 0.05)(cid:18) σpub
1000km/s(cid:19)1.06±0.08
(30)
with a dispersion of 0.16 dex that is consistent with the
uncertainties.
If we restrict the fit to the 69 systems
with Nv ≥ 10 redshift measurements, we find essentially
the same parameters. Hence our membership probability
weighting method for determining velocity dispersions is
consistent with the more standard methods based on re-
jecting outliers.
14
1000
100
100
1000
Fig. 8. -- Comparisons between our velocity dispersion estimates
and published estimates. Filled points have Nv ≥ 10 velocities and
open points have 5 ≤ Nv < 10 velocities. The dashed lines show
the best fit relation and its width, which would lie on the solid line
if the agreement was perfect.
We also find 173 (108) clusters with X-ray luminosities
(temperatures) and at least Nv ≥ 5 galaxies with redshifts
from from the sources summarized in §3.6. Figs. 9 and
10 show our estimates of the relations between the X-ray
luminosity LX and the X-ray gas temperature TX to the
velocity dispersion estimates. The correlation between the
velocity dispersion and the X-ray luminosity, including er-
rors in both quantities, is
log L44 = (0.20 ± 0.07) + (4.47 ± 0.24) log(cid:18)
σ
103km/s(cid:19)
(31)
where L44 = LX h2/1044ergs/s. The variance in the X-ray
luminosity at fixed velocity dispersion is a (typically) large
0.76 dex. If we restrict the fit to systems with Nv ≥ 10 the
normalization of 0.18± 0.08 is little changed, but the slope
steepens to 4.69±0.28 with a reduced variance of 0.55 dex.
The slope and normalization are close to those found by
Mulchaey & Zabludoff (1998, zero point 0.48 ± 1.09, slope
4.29±0.37) or Mahdavi & Geller (2001, zero point 0.40+0.9
−2.0,
slope 4.4+0.7
−0.3). Wu et al. (1999) found a range of slopes
from 2.56 ± 0.21 to 5.24 ± 0.29 depending on their fit-
ting methodology. The zero-point uncertainties we cite
for these published relations are overestimates because we
had to correct the relations to a velocity scale 103 km/s
from one of 1 km/s without knowing the full covariance
matrix of the fit. For the 108 clusters with X-ray temper-
atures, we find
log(cid:18)
σ
103km/s(cid:19) = (0.01±0.02)+(0.63±0.04) log(cid:18) TX
6keV(cid:19)(32)
with a spread of 0.16 dex, which is close to the rela-
1000
100
1000
100
0.001
0.01
0.1
1
10
0.1
1
10
Fig. 9. -- The velocity dispersion estimate σ as a function of the
X-ray luminosity LX . Filled points have Nv ≥ 10 velocities and
open points have 5 ≤ Nv < 10 velocities. The light solid line is the
correlation from Mulchaey & Zabludoff (1998) and the heavy solid
line is the correlation from Mahdavi & Geller (2001).
Fig. 10. -- The velocity dispersion estimate σ as a function of the
X-ray temperature TX . Filled points have Nv ≥ 10 velocities and
open points have 5 ≤ Nv < 10 velocities. The light solid line is the
fit from Wu, Fang & Xu (1998) with slope 0.56, the heavy solid line
the fit with slope 0.67. The heavy dashed line is the estimate from
Girardi et al. (1998).
tions found by Mulchaey & Zabludoff (1998, zero point
−0.01±0.04, slope 0.51±0.05), Wu et al. (1999, zero point
−0.00 ± 0.03, slope 0.65 ± 0.02), and Girardi et al. (1998,
zero point −0.01 ± 0.03, slope 0.62 ± 0.04). Fitting to
the 79 systems with Nv ≥ 10 gives the same parameters
(0.01 ± 0.02 and 0.62 ± 0.04) with a reduced scatter of
0.12 dex. The similarity with these published analyses sug-
gests that our method for determining velocity dispersions
is comparable to those used for other studies despite the
small numbers of galaxies used in our dispersion estimates
compared to these more extensive surveys of individual
clusters.
5.3. The Halo Multiplicity Function
We now derive the halo multiplicity function based on
the three observables, velocity dispersion, X-ray luminos-
ity and X-ray temperature, which we can later relate to
the cluster mass. We first derive the observed correlations
and then estimate the conversion to the desired correlation
with cluster mass.
Fig. 11 shows that the observed relationship between
velocity dispersion and galaxy number is very similar to
that in the synthetic data (§4, Fig. 5). The best fit relation
for the 939 systems with Nv ≥ 5, including the errors in
both quantities, is
log N∗666 = (1.37 ± 0.03) + (2.63 ± 0.06) log(σ/103km/s)
(33)
compared to 1.32± 0.04 and 2.76± 0.08 for the zero point
and slope of the fits to the synthetic catalog (1.41 and 2.75
for the input relation). Only one system is dropped from
the fits. The scatter in N∗666 at fixed dispersion is slightly
larger in the real data (0.53 dex versus 0.50 dex). If we
restrict the fit to the 238 systems with Nv ≥ 10, we obtain
consistent parameters (1.31± 0.04 and 2.60± 0.10) with a
reduced scatter of 0.41 dex. While the significance of the
15
fit in Eq. (33) is quite high, the very large scatter makes it
hard to see by eye. Therefore as a check we also used the
dispersion in the distribution of galaxy-group velocities in
'stacked' clusters, binned in 5 logarithmic bins of N∗666,
to make the trend more clearly visible. We obtained a fit
very close to Eq. (33).
Next we compare the halo occupation number to the
X-ray properties of the clusters, as shown in Figs. 12 and
13. We must face two problems in correlating N∗666 with
the X-ray data. First, we find that the scatter in the cor-
relations is significantly larger than is consistent with the
uncertainties in either quantity. In part this is due to the
presence of significant intrinsic scatter in the X-ray proper-
ties of clusters. Second, there is considerable evidence for
a break in the X-ray properties between groups and clus-
ters, generally believed to be due to significant cooling and
heating from star formation processes in the lower mass
systems (David et al. 1993; Ponman et al. 1996; White,
Jones & Forman 1997; Allen & Fabian 1998; Markevitch
1998; Arnaud & Evrard 1999; Helsdon & Ponman 2000;
Finoguenov, Reiprich & Bohringer 2001). Between the in-
trinsic scatter and the break in the properties, the param-
eters for power-law fits to X-ray correlations can depend
on the fitting methods and the data used.
In order to include the effects of these problems we
made two modifications to our fitting procedures. First,
we added a logarithmic systematic error (or intrinsic scat-
ter) σsys in quadrature with the measurement errors for
the fitted variables. We then adjusted σsys so that the fit
has χ2 = Ndof. Second, we fit power-law relations includ-
ing the X-ray data only for the more massive clusters with
TX ≥ 1 keV and LX ≥ 1042h2 ergs/s. This reduces the
biases in the correlations from attempting to fit the lower
mass systems where non-adiabatic effects begin to modify
the X-ray properties while leaving enough systems to ob-
10
1
0.1
100
10
1
0.1
100
1000
0.01
0.1
1
10
Fig. 11. -- The number of galaxies N∗666 as a function of the
cluster velocity dispersion σ for the 2MASS data. Points and lines
as in Fig. 5.
Fig. 12. -- The number of galaxies N∗ versus the X-ray tem-
perature TX . The dashed lines show the best power law fit for
TX ≥ 1 keV and the average dispersion of the data about the fit.
16
tain a statistically significant fit. As in all our standard
fits we include only the systems with Nv ≥ 5 associated
redshifts.
If we now find the best power-law relation between LX
and TX, we find
log L44 = (0.21 ± 0.04) + (2.66 ± 0.12) log(cid:18) T
6keV(cid:19) (34)
based on 84 clusters with σsys = 0.08 and a mean scatter of
0.30 dex. The normalization is consistent with our adopted
standard from Markevitch (1998) in Eq. (18), but it has a
significantly steeper slope. If we include no limits on the
X-ray properties we find a similar zero-point of 0.28± 0.05
but a still steeper slope of 3.04 ± 0.10 and a larger scatter
of 0.35 dex. This fit matches that of Xue & Wu (2000)
under similar assumptions. There is clearly a break in the
slope of the LX -- TX relation near TX = 1 keV which is
incompatible with a single power law providing a good fit.
We first fit the number-temperature relation as a power
law using 84 clusters, as shown in Fig. 12, to find that
log N∗666 = (1.38±0.05)+(2.09±0.17) log(T /6keV). (35)
with σsys = 0.12 dex and a mean scatter of 0.35 dex. This
relationship changes little if we drop the restrictions on the
X-ray properties. The fit for all 102 clusters with X-ray
temperatures, luminosities and Nv ≥ 5, has a zero-point of
1.39± 0.05, a slope of 1.97± 0.12, and a negligible increase
in σsys and the scatter.
Fig. 13 shows the results for fitting the 153 clusters with
LX ≥ 1042h−2 ergs/s and Nv ≥ 5. We find
log N∗666 = (1.29 ± 0.05) + (0.75 ± 0.05) log L44.
(36)
100
10
1
0.1
0.01
0.001
0.01
0.1
1
10
Fig. 13. -- The number of galaxies N∗ versus the X-ray luminosity
LX . The dashed curves show our best power-law fit to the relation
for LX ≥ 1042h2 ergs/s. The solid curves show our theoretically
estimated conversion (Eq. 22) of the Λcl -- LX correlation found by
Donahue et al. (2001) and the uncertainties in its normalization.
We can match the two relations by raising the normalization of the
conversion to Λcl ≃ 21N
3/4
∗666.
for σsys = 0.24 and with a scatter of 0.33 dex. If we in-
clude all 173 clusters with X-ray luminosity measurements
and Nv ≥ 5, the zero-point becomes 1.23 ± 0.04 and the
slope of 0.64±0.03 is significantly flatter. The distribution
of clusters in Fig. 13 is clearly inconsistent with a simple
power law if we extend the fit to low X-ray luminosities,
so we will not consider this case further. As a consistency
check, we can combine the N∗666 -- TX and N∗666 -- LX rela-
tions (Eqs. 36 and 35) to infer a zero-point of 0.12 ± 0.09
and a slope of 2.79 ± 0.29 for the LX -- TX relation which
are consistent with the direct fit in Eq. (34). Similarly, if
we combine the N∗666 -- σ and N∗666 -- TX relations (Eqs. 33
and 35), we infer a zero-point of 0.00 ± 0.02 and a slope
of 0.79 ± 0.07 for the σ -- TX relation which are reasonably
consistent with the direct fit in Eq. (32).
We can also compare our results to those of Donahue et
al. (2001), who compared the properties of clusters found
in X-ray surveys with those found using the Postman et
al. (1996) matched filter algorithm. In this algorithm, clus-
ters are characterized by their total luminosity in units of
L∗, Λcl, where we estimated a conversion of Λcl = 11N 0.75
∗666
(see Eq. 22) between our cluster normalizations. Given
this relation, we can convert their estimated correlation
with X-ray luminosity into our units,
log N∗666 = (1.8 ± 1.0) + (0.8 ± 0.2) log L44.
(37)
While the slope and the normalization are consistent with
our estimate given the uncertainties in their relation, the
normalization is systematically higher.
If we estimate
the conversion between Λcl and N∗666 by matching the
X-ray correlations for the two variables, we find that
Λcl = 21N 0.83
∗666, where the slope is consistent with our the-
oretical estimate (Eq. 22) but the normalization is nearly
doubled. Since the sense of the change is the same as
that needed to better match Λcl and N∗666 for clusters of
the same Abell richness class (see §3.6), we will adopt an
empirical relation for the conversion of
Λcl ≃ 21N 3/4
∗666
(38)
for subsequent comparisons. Using the original LX and Λcl
data points, we confirmed that this empirical conversion
was correct, while the theoretical conversion from Eq. (22)
showed an offset whose origin we do not fully understand.
The final step in the analysis is to convert the ob-
served occupancy functions (N∗666(σ), N∗666(TX ) and
N∗666(LX )) into an estimate of the occupancy function
N∗666(M200) as a function of the virial mass. The results,
summarized in Table 2, are dominated by systematic errors
associated with the choice of the mass scale. The average
normalization A = 1.25 ± 0.03 and slope B = 1.02 ± 0.10
found when we set the mass scale using the velocity dis-
persion (σ(M200), Eq. 28) are lower than the average nor-
malization A = 1.58±0.02 and slope B = 1.21±0.14 found
when we set the mass scale using the X-ray temperatures
(M200(TX ), Eq. 19). The results based on the same mass
scale are largely independent of the observed occupancy
function used to make the estimate. This is particularly
true of the normalization A, where the scatter between the
individual estimates is less than the formal errors, and less
true of the slope B where the scatter between the individ-
ual estimates is somewhat larger than the formal errors.
A shallower mass-temperature relation with T ∼ M 1/2
rather than T ∼ M 2/3, perhaps better suited to the mix
of high and low temperature clusters we use, would bring
the two estimates of the slopes into agreement. In addition
to these systematic differences, recall that in the synthetic
data our fits to the occupancy function underestimated the
normalization by (19 ± 7)% and the slope by 0.06 ± 0.03
(see §4.2 and Table 2).
5.4. The Cluster Number Function
Next we estimate the cluster number function, dn/dN∗666,
for the 2MASS sample as shown in Fig. 14. The number
function for the real sample extends to higher values of
N∗666 than that for the synthetic sample (see Fig. 6) but
otherwise looks very similar. The results for the three val-
ues of Nthresh are again mutually consistent. Note that for
large N∗666 the number function is systematically above
that of the synthetic catalog.
Comparing our results with previous estimates of the
same quantity is quite difficult, due to different conven-
tions for estimating 'richness'. The best we can do is to ap-
ply reasonable transformations based on scaling relations
between the widely differing definitions, realizing that this
method is far from perfect.
Donahue et al. (2001) and Postman et al. (2002) have
estimated the density of clusters at intermediate redshift
0.1
1
10
100
Fig. 14. -- The number function dn/dN∗666 for the 2MASS sam-
ple. The open triangles, filled squares and open pentagons show the
results for Nthresh = 3, 5 and 10. In many cases these points overlap.
The bootstrap error estimates are shown only for Nthresh = 5. The
smooth curve is the number function expected for the synthetic cat-
alog in Fig. 6. The heavy solid line and the filled triangles show the
dn/dΛcl number functions from Donahue et al. (2001) and Postman
et al. (2002) respectively, converted to our units using the empirical
conversion of Eq. (38). The lower dashed curve cutting off sharply
at high mass shows the luminosity function from the NOG in Mari-
noni et al. (2000) converted to a number function as described in
the text. The other dashed curve shows the fit to the CfA survey
luminosity function from Moore, Frenk & White (1993) converted
in the same way. The equivalent mass scale shown at the top of the
figure is the conversion from N∗666 to M200 estimated in Eq. 41.
in terms of Λcl (see §3.6 and Eq. 20) over the ranges 20 <∼
Λcl <∼ 100 and Λcl >∼ 30 respectively. They fit the number
density as a power law of the form
17
(39)
dn
dΛcl
40(cid:19)−α
= N0(cid:18) Λcl
75Mpc−3 with α = 5.3± 0.5 and
finding N0 = 6+3
−1 × 10−6h3
75Mpc−3 with α = 4.40 ± 0.30
N0 = (1.55 ± 0.40) × 10−6h3
respectively. If we use the empirical conversion between
Λcl and N∗666 derived from comparing the two richness
estimates at fixed X-ray luminosity (Eq. 38), then we find
good agreement for the number density of N∗666 ∼ 10
If, however, we use our theoreti-
clusters (see Fig. 6).
cal estimate of the conversion (see §3.6, Eq. 22), then the
converted Donahue et al. (2001) and Postman et al. (2002)
number densities are nearly an order of magnitude higher
than our own estimates. Since both our catalogs (§5.1)
and Donahue et al. (2001, 2002) have similar complete-
ness compared to X-ray surveys, the need to use the em-
pirical conversion in order to obtain similar number den-
sities suggests that there is a problem with the theoretical
conversion between the two richness estimates.
Marinoni, Hudson & Giuricin (2000) used the Nearby
Optical Galaxy (NOG, Marinoni 2001) sample to make a
similar estimate based on the group catalogs of Giuricin
et al. (2001). If ΛG = Ltot/L∗ is the estimate for the total
group luminosity Ltot in units of the luminosity of an L∗
galaxy, they derive a number function of
n0
=
(40)
dn
dΛG
ΛG0(cid:18) ΛG
ΛG0(cid:19)−αG
exp(−ΛG/ΛG0)
75 × 10−4 Mpc−3, and
where αG ≃ −1.4, n0 = 4.8h3
ΛG0 ≃ 9.9. For a galaxy luminosity function with their
estimated Schechter slope of α = −1.1, the equivalent
number of L > L∗ galaxies is NG0 ≃ ΛG0/5.1 = 1.9. For
groups selected with a standard FoF algorithm, we expect
N∗666 ≃ 0.66NG, but the applicability of this conversion
to the Marinoni et al. (2000) estimates is just a rough esti-
mate. The number function dn/dN∗666 is identical to the
Schechter function dn/dΛG with ΛG replaced by N∗666 and
ΛG0 = 1.3. This estimate, which is also shown in Fig. 14,
is a factor of 5 higher than ours for low N∗666 and cuts
off more sharply. The sharper cutoff is simply a conse-
quence of the cz < 6000 km/s redshift cutoff of the NOG
Sample, which eliminates all the nearby rich clusters. The
normalization could be brought into agreement with our
estimate if we have overestimated the coefficient in the
relation N∗666 ≃ 0.66NG used to convert the FoF mem-
bership into our standard overdensity. Finally we have
converted the luminosity function of the CfA sample, as
derived by Moore, Frenk & White (1993), to dn/dN∗666
in the same manner as described above and with the same
difficulties. This is plotted as the other dashed curve which
agrees with the NOG estimate at small N∗666 but has sig-
nificantly more objects at high N∗666. As with the compar-
ison to Donahue et al. (2001) and Postman et al. (2002),
differences in the absolute number density are very sensi-
tive to errors in the conversion between richness estimates,
and we lack any means of doing this more precisely given
the information available in Marinoni et al. (2000) and
Moore et al. (1993).
18
Table 2 also gives the estimates for the halo occupancy
function needed to transform the Jenkins et al. (2001)
mass function for our assumed cosmology into the ob-
served number function (as in section 4.3). The results for
Nthresh = 3, 5 and 10 are mutually consistent and inter-
mediate to the results from fitting the observed occupancy
functions in §5.3.
5.5. Derived Quantities
Finally we discuss the implied infrared mass-to-light ra-
tio of clusters and the fraction of the local luminosity den-
sity associated with the virialized regions of clusters. We
start by adopting a standard model for the halo occupa-
tion function. Our uncertainties in the occupation func-
tion are dominated by systematic errors in how to relate
the observed quantities to cluster masses. Depending on
whether we set the mass scale using the velocity disper-
sion, the X-ray temperature, or the cluster mass function,
we find significant differences in both the normalization
and the slope of the occupancy function (see Table 2). We
decided to simply average the results for the three possi-
ble mass scales and use their dispersions as the standard
errors:
log N∗666 = (1.44±0.17)+(1.10±0.09) log(M200h/1015M⊙).
(41)
The resulting errors are larger than the statistical un-
certainties for any given mass normalization method and
somewhat larger than the systematic offsets we found
when estimating the occupancy function of the synthetic
catalog. The normalization is higher than we used in the
synthetic catalog, which seems reasonable given the visibly
weaker fingers of god in the synthetic catalogs.
0.1
1
10
100
Fig. 15. -- The fraction jclust(> N∗666)/jtotal of all stellar lumi-
nosity in systems with at least N∗666 galaxies. The open triangles,
filled squares and open pentagons show the results for Nthresh = 3,
5 and 10. In many cases these points overlap. The bootstrap error
bars are shown only for Nthresh = 5. The equivalent mass scale
shown at the top of the figure is the conversion from N∗666 to M200
estimated in Eq. 41.
From Eq. (41) we can estimate more traditional quan-
tities like the K-band mass-to-light ratios of the clusters.
Including the uncertainties in the luminosity function, the
correction from aperture to total magnitudes, and a zero
point of M⊙Ks = 3.39 mag (see Kochanek et al. 2001 and
Cole et al. 2001) we find that
L666(cid:19)K
(cid:18) M200
= (116 ± 46)h(cid:18) M200h
1015M⊙(cid:19)0.10±0.09
(42)
where almost all the uncertainty comes from the uncer-
tainty in the normalization A = 1.44 ± 0.17 of the oc-
cupancy function (it would be (116 ± 10)h for A ≡ 1.44).
This is nearly identical to the baseline semi-analytic model
of Cole et al. (2000) where the K-band mass-to-light ra-
tio is 118(M200h/1015M⊙)0.06 for clusters with masses
above 1013.5h−1M⊙. The mass-to-light ratio is essen-
tially independent of mass, which differs somewhat from
estimates in the B-band (e.g. Girardi et al. 2000, Mari-
noni & Hudson 2001) which suggest a steeper slope of
(M/L)B ∝ M 0.1−0.3, albeit with significant uncertain-
If there is a real slope difference,
ties in the slopes.
with (M/L)B/(M/L)K ∝ M x then we can explain this
as a trend in the colors of the galaxy population with
∆(B − K) = 5x ∼ 0.5 to 1.5 mag over the range from
1013M⊙ to 1015M⊙. All general trends in cluster prop-
erties have the sense required to produce the slope dif-
ference. Starting from a constant K-band mass-to-light
ratio, the more massive clusters could be dimmer in the
B-band because of the increasing early-type galaxy frac-
tion (a B -- K color shift of roughly 0.5 mag from Sa to E),
increasing metallicity (∆(B -- K)/[F e/H] ∼ 1.2 mag/dex)
or increasing age (∆(B -- K)/ log t ∼ 0.5 to 1.0 mag/dex).
Like the halo occupation number itself, it is important
to match definitions when comparing our mass-to-light
ratios to other estimates. Our estimate is normalized
to be the ratios of the two quantities in spheres of ra-
dius rN 666 ≃ rM200, leading to higher mass to light ra-
tios than estimates comparing the light in cylinders to
the mass in spheres. For example, Rines et al. (2001)
found (M/L)K = (75 ± 23)h for Coma using the 2MASS
survey, and the Kochanek et al. (2001) luminosity func-
tion. At the rN 666 = 1.5h−1 Mpc, which also matches the
Rines et al. (2001) estimate of the virial radius, we must
raise their estimate of the mass to light ratio by 25% to
(M/L)K = (92 ± 28)h to compare it to our estimates of
the mass-to-light in spheres. This is quite close to our esti-
mate of (M/L)K = (116± 46)h for the mass-to-light ratio
of 1015h−1M⊙ cluster like Coma in Eq. (42).
Since the K-band luminosity is well correlated with the
total stellar mass, we can use the cluster number function
to obtain a rough estimate of the fraction of stars in sys-
tems above a given mass or number threshold. The total
luminosity density of all galaxies is jtot = n∗L∗Γ[2 + α]
while the equivalent luminosity density contained clusters
with N∗666 > N (and in the region with r < rN 666) is
jclust(> N ) =
L∗Γ[2 + α]
Γ[1 + α, 1] Z ∞
N
dN∗666N∗666
dn
dN∗666
. (43)
The ratio jclust/jtot is shown in Fig. 15 as a function of
N∗666. We find that the virialized regions of clusters more
massive than 1014 h−1M⊙ contain ≃ 7% of the local lumi-
nosity. This is quite comparable to the fraction of the mass
contained in such clusters in popular ΛCDM models with
Ωmat ≃ 0.3 and σ8 ≃ 1, though the number is particularly
sensitive to the latter assumption. The enclosed fraction
also depends on the density contrast to which we extend
the cluster edges. For example, if we include galaxies out
to ∆N = 100, the luminosity fraction in clusters doubles.
Since our data is consistent with N (M ) ∼ M we would
predict that 2MASS galaxies provide good tracers of the
mass with a small and relatively scale-independent bias on
Mpc scales. This would suggest that the correlation func-
tion (or power spectrum) of 2MASS galaxies should show
the virial 'inflection' of the dark matter power spectrum
on Mpc scales, where the growth of fluctuations is first en-
hanced by non-linear infall and then suppressed by virial
motions within halos. There is weak evidence for this in
Fig. 3 of Allgood, Blumenthal & Primack (2001), though
further work with the full catalog needs to be done to re-
ally test this prediction. That N (M ) ∼ M is perhaps not
unexpected, as both simulations and observations show
that galaxies selected in the "red" trace the matter dis-
tribution better than those selected in the "blue". The
original semi-analytic models (Kauffman et al. 1999) pre-
dicted a steeper slope for N (M ) for their red galaxies, with
N (M ) ∼ M 0.9, which is consistent with our results.
Theory predicts, and our simulations have assumed,
that the number of galaxies in a cluster mass halo should
follow a Poisson distribution (at fixed mass).
It is diffi-
cult to test this with our current sample due to the dif-
ficulty of estimating the mass of our clusters. The ratio
of hN (N − 1)i1/2/hNi, where N is the number of mem-
ber galaxies determined from the matched filter probabil-
ities pi, is very close to unity for clusters binned by N∗666.
(The Poisson prediction is hN (N − 1)i1/2 = hNi.) How-
ever N∗666 is not a good proxy for M200, having significant
scatter. For the small number (110) of clusters for which
we can detect all L∗ galaxies and we have an X-ray tem-
perature, the ratio hN (N − 1)i1/2/hNi is also consistent
with unity, though we need to choose broad bins in TX
and the statistics are poor.
6. CONCLUSIONS
We have applied the matched filter technique to both
simulated galaxy catalogs and the 2MASS galaxy cata-
logue to search for clusters over approximately 90% of the
sky to a redshift limit of z ≃ 0.05. We have matched our
2MASS derived catalog to existing catalogs in an auto-
mated way using the NASA Extragalactic Database. Our
algorithm appears to be both robust and efficient, return-
ing quite complete samples of clusters out to distances
where the typical cluster contains as few as 3 galaxies. The
algorithm finds almost no 'false groups', the main source
of contamination being the inclusion of objects below the
mass cut in the catalog.
The matched filter algorithm gives us a way of estimat-
ing cluster membership and hence cluster properties like
the multiplicity function, the number function and the ve-
locity dispersion in a new way. While our estimates for the
velocity dispersion agree well with earlier work, we typi-
cally have far fewer galaxies per cluster than dedicated
surveys and hence larger errors. This is offset by the fact
that we have a large number of clusters, over much of the
sky, with which to search for correlations between cluster
properties such as velocity dispersion and X-ray tempera-
19
ture. Where there is overlap we find quite good agreement
with earlier work, though often with improved statistics.
Although it is necessary to be very careful about the
definition of N in order to compare with theory, we find
that with sufficient care we have been able to estimate the
cluster number function, dn/dN , over more than 2 orders
of magnitude in N .
By using the velocity dispersion, X-ray luminosity or
X-ray temperature as a surrogate for mass we are able to
estimate the multiplicity function N (M ). Although there
are serious issues in converting observables into cluster
masses, all of our results are fairly consistent and suggest
that N (M ) ∼ M or perhaps slightly steeper, in reasonable
agreement with earlier estimates. With all quantities nor-
malized by the spherical radius corresponding to a mass
overdensity of ∆M = 200 or the equivalent galaxy num-
ber overdensity of ∆N = 200Ω−1
M = 666, we find that the
number of L > L∗ galaxies in a cluster of mass M200 is
log N∗666 = (1.44±0.17)+(1.10±0.09) log(M200h/1015M⊙).
(44)
The uncertainties in this relation are largely due to the
choice made for relating the observed quantities to the
cluster mass scale. For a fixed mass scale the scatter
resulting from the different observed correlations is con-
siderably smaller. Correlations of N with other cluster
properties, X-ray luminosity, temperature or galaxy ve-
locity dispersion, are given in §5.3. The region has a K-
band cluster mass-to-light ratio of (M/L)K = (116± 46)h
which is essentially independent of cluster mass. The un-
certainties are again dominated by the choice of the mass
scale. This scaling is consistent with N (M ) ∼ M , though
if we take our best fit seriously and N (M ) is steeper
than M we expect M/L would fall slowly with increas-
ing mass. Integrating over all clusters more massive than
M200 = 1014 h−1M⊙, the virialized regions of clusters con-
tain 7% of the local stellar luminosity, quite comparable
to the (somewhat theory dependent) mass fraction in such
objects in currently popular ΛCDM models.
The cluster likelihoods tend to be larger in the real data
than our mock catalogs. This difference could have sev-
eral sources. It could be due to the concentration of red-
shift measurements in the real catalog towards groups and
clusters rather than having the random distribution of the
synthetic catalog. It may also indicate that the cosmol-
ogy adopted in the underlying simulation is incorrect (the
number of clusters is very sensitive to σ8 for example),
the normalization of the input N (M ) may be too low,
our assumed N (M ) may have too many galaxies in low-
mass halos compared to high-mass halos, or the luminos-
ity function could vary systematically with the parent halo
mass rather than being fixed (as we have assumed). Along
these lines, both the expectation that light closely trace
mass in K-band and that the luminosity function change
with mass suggest that clusters contain a larger fraction of
galaxies than we have assumed in our current generation
of mock catalogs. These avenues will be explored in the
future with larger simulations and better data.
We view this work as but the first step in an iterative
sequence. Based on simulations which we know describe
reality imperfectly, we have calibrated our cluster finding
algorithm. When applied to the real data this algorithm
allows us to estimate correlations between different prop-
20
erties of a halo which can be used in the next stage to
improve the simulated catalogs. As the 2MASS catalog
becomes increasingly complete, and more redshifts become
available, we will estimate the global relations between dif-
ferent halo parameters by providing them as priors to the
fitting, and varying the parameters to optimize the global
likelihood. These relations can then be used in the con-
struction of the galaxy catalogs from improved simulations
which will allow us to further optimize and understand the
cluster finding algorithm.
ACKNOWLEDGMENTS
We would like to thank the 2MASS collaboration for
their cooperation and comments on this paper. We would
also like to thank N. Caldwell, M. Donahue, D. Eisen-
stein, C. Jones, M. Pahre, M. Postman and K. Rines for
helpful conversations and C. Baugh & S. Cole for pro-
viding additional IR properties for clusters in their semi-
analytic models. M.W. would like to thank H. Mo and
R. Yan for enlightening conversations on the halo model
of galaxy clustering. 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 Insti-
tute of Technology, funded by the National Aeronautics
and Space Administration and the National Science Foun-
dation. This research has made use of the NASA/IPAC
Extragalactic Database (NED) which is operated by the
Jet Propulsion Laboratory, California Institute of Tech-
nology, under contract with the National Aeronautics and
Space Administration. C.S.K. and J.P.H. were supported
by the Smithsonian Institution. M.W. was supported by
a Sloan Fellowship, the NSF and NASA. Simulations were
carried out at CPAC, through grants PHY-9507695, and at
the National Energy Research Scientific Computing Cen-
ter.
APPENDIX
INCLUDING COLOR INFORMATION
In this appendix we outline the inclusion of color in-
formation in our algorithm. To include color information
in our algorithm we require a model for the multi-variate
luminosity function in luminosity and colors, φ(M, C0),
and its evolution with redshift. The multi-variate and
standard luminosity functions are related by φ(M ) =
R φ(M, C0)dC0. Assuming that the numbers of galaxies
are not evolving, the measured magnitudes and colors, m
and c are related to the local values by M = m − D(z)
(Eq. 1) and C0 = c − C(M, z) to remove the effects of
distance, K-corrections and evolution.
For the 2MASS survey the description simplifies fur-
ther because only the H -- K color has a significant cor-
relation with redshift and neither the H -- K nor the J -- H
color has a significant correlation with luminosity. This
means that we can factor the bivariate luminosity func-
tion, φ(M, C0) = φ(M )ξ(C0), with R ξ(C0)dC0 ≡ 1. The
color distribution, which consists of a narrow peak with a
red tail, is well modeled by the sum of two Gaussians,
ξ(H−K) =
f
√2πσ0
e−(H−K−a0)/2σ2
0 +
1 − f
√2πσ1
e−(H−K−a1)/2σ2
1
(A1)
where f = 0.698, a0 = 0.226, a1 = 0.268, σ0 = 0.025 and
σ1 = 0.081. The galaxies become steadily redder at higher
redshift, with
C(z) = 2.333z − 0.770z2,
(A2)
and it appears broader at fainter magnitudes due to mea-
surement errors that can be modeled by convolving the dis-
tribution with a Gaussian of width σ = 0.092×100.2(K−13).
While the errors in the flux are unimportant to our analy-
sis given the sample magnitude limit, the shallow slope of
C(z) means that we must take into account the errors in
the H -- K colors.
Given the bivariate luminosity function and the depen-
dence of the colors on redshift, we can modify the terms of
the likelihood function (see §3.2 and §3.3) to include the
additional information. The probability of finding a field
galaxy with magnitude mi, color ci and redshift zi is
dDc
dz
φf (m−D(zi))ξ(ci−C(zi)).
Pf (mi, ci, zi) = 0.4 ln 10D2
C(zi)
(A3)
Integrating over redshift, we obtain the number counts
of galaxies with a given color, which is equivalent to the
probability of finding a galaxy with magnitude mi and
color ci,
Pf (mi, ci) =Z ∞
0
Pf (mi, ci, z)dz.
(A4)
The probability distribution for an unknown galaxy red-
shift is simply the ratio Pf (mi, ci, zi)/Pf (mi, ci). Because
the slope of C is shallow, the addition of the H -- K color in-
formation only modestly improves the accuracy of redshift
estimates. For the 50000 galaxies with known redshifts,
the observed errors in the predicted redshifts are mod-
estly smaller than their theoretical estimates of σz = 0.028
(using only fluxes), 0.036 (using only colors) and 0.024
(using both). The errors in redshifts estimated using col-
ors (fluxes) grow slowly (linearly) with redshift, so at low
redshifts fluxes always become better redshift estimators
than colors. We must also modify the cluster membership
probabilities, replacing φc(M ) by φc(M, C0), while leav-
ing the normalization factor in the denominator, Φc(M ),
unchanged.
This formalism for including color information in a
matched filter algorithm can be generalized to additional
colors and more complicated descriptions for the evolution
of colors with redshift and luminosity, although tabulated
models will probably be required. Matched filters can also
exploit the redder optical colors of cluster galaxies (as used
in other algorithms, e.g. Goto et al. 2001) by using a multi-
variate cluster luminosity function with a different color
distribution from that for the field.
REFERENCES
Abell G.O., 1958, ApJS, 3, 211
Allen S.W., Fabian A.C., 1998, MNRAS, 297, 63
Allgood B., Blumenthal G., Primack J.R., 2001,
in Proceedings
of International School of Space Science 2001, ed. Aldo Morselli
(Frascati Physics Series) v2 [astro-ph/0109403]
Arnaud M., Evrard A.E., 1999, MNRAS, 305, 631
Bahcall N., et al., preprint [astro-ph/0205490]
Barmby, P., & Huchra, J.P., 1998, AJ, 115, 6
Bartelmann, M., 1996, A&A, 313, 697
Berlind, A.A., & Weinberg, D.H., 2002, ApJ in press [astro-
ph/0109001]
Berlind, A.A., et al., 2002, in preparation.
Blanchard A., Sadat R., Bartlett J.G., le Dour M., 2000, a, 362, 809
Bohringer H., et al., 2001, A&A, 369, 826
Bohringer H., et al., 2000, ApJS, 129, 435
Borgani S., Girardi M., Carlberg R., Yee H., Ellingson E., 1999, ApJ,
Carlberg R., et al., 1996, ApJ, 462, 32
Carlstrom J.E., et al., 2000, Phys. Scripta, 85, 148
Chiu, W.A., Gnedin, N.Y., Ostriker, J.P., 2001, preprint [astro-
527, 561
ph/0103359]
Kauffmann G., Colberg J.M., Diaferio A., White S.D.M., 1999,
Katz N., Hernquist L., Weinberg D.H., 1999, ApJ, 523, 463
Kepner, J., Fan, X., Bahcall, N., Gunn, J., Lupton, R., & Xu, G.,
MNRAS, 303, 188
1999, ApJ, 517, 78
Kim, R.S.J., et al., astro-ph/0110259
Kochanek, C.S., 2001,
in the proceedings of The Dark Universe
meeting at STScI, April 2-5, 2001 M. Livio, ed., Cambridge
University Press
Kochanek, C.S., Pahre, M.A., Falco, E.E., Huchra, J.P., Mader, J.,
Jarrett, T.H., Chester, T., Cutri, R., & Schneider, S.E., 2001, ApJ,
560, 566
Lewis A.D., Stocke J.T., Ellingson E., Gaidos E.J., 2001, preprint
[astro-ph/0110156]
Lin, H., Kirshner, R.P., Shectman, S.A., Landy, S.D., Oemler, A.,
Christlein, D., 2000, ApJ, 545, 145
Cole, S., Lacey, C.G., Baugh, C.M., Frenk, C.S., 2000, MNRAS, 319,
Tucker, D.L., & Schechter, P.L., 1996, ApJ, 464, 60
Lumsden S.L., Nichol R.C., Collins C.A., Guzzo L., 1992, MNRAS,
21
168
Cole, S., Norberg, P., Baugh, C.M., et al., 2001, MNRAS, 326, 255
Cruddace, R., Voges, W., Boehringer, H., Collins, C.A., Romer,
[astro-
A.K., MacGillivray, H., Ye, 2001, ApJS,
ph/0201069]
in press
Dalcanton J., 1996, ApJ, 466, 92
Dalton G.B., Efstathiou G., Maddox S.J., Sutherland W.J., 1992,
David, L.P., Slyz, A., Jones, C., Forman, W., Vrtilek, S.D., Arnaud,
ApJ, 390, L1
K.A., 1993, ApJ, 412, 479
De Grandi, S., Bohringer, H., Guzzo, L., Molendi, S., Chincarini, G.,
Collins, C., Cruddace, R., Neumann, D., Schindler, S., Schuecker,
P., & Voges, W., 1999, ApJ, 514, 148
Diaferio, A., Kauffmann, G., Colberg, J.M., & White, S.D.M., 1999,
MNRAS, 307, 537
Donahue, M., Mack, J., Scharf, C., Lee, P., Postman, M., Rosati, P.,
Dickinson, M., Voit, G.M., & Stocke, J.T., 2001, ApJL, 552, L93
Donahue, M., Scharf, C., Mack, J., Lee, J.P., Postman, M., Rosati,
P., Dickinson, M., Voit, G.M., & Stocke, J.T., 2002, ApJ, 569, 689
Dressler, A., 1980, ApJ, 236, 351
Ebeling, H., Voges, W., Bohringer, H., Edge, A.C., Huchra, J.P., &
Briel, U.G., 1996, MNRAS, 281, 799
Ebeling H., Edge, A.C., Bohringer, H., Allen, S.W., Crawford, C.S.,
Fabian, A.C., Voges,W., & Huchra, J.P., 1998, MNRAS, 301, 881
Ebeling H., Edge, A.C., Allen, S.W., Crawford, C.S., Fabian, A.C.,
& Huchra, J.P., 2000, MNRAS, 318, 333
Ebeling H., Edge A.C., Henry P., 2001, preprint [astro-ph/0009101]
Edge A., Stewart G.C., Fabian A.C., Arnaud K.A., 1990, MNRAS,
245, 559
Evrard A.E., Metzler C.A., Navarro J.F., 1996, ApJ, 469, 494
Evrard A.E., et al., 2001, preprint [astro-ph/0110246]
Finoguenov A., Reiprich T.H., Bohringer H., 2001, A&A, 368, 749
Frederic, J.J., 1995, ApJ, 97, 259
Gardner, J.P., Katz, N., Hernquist, L., Weinberg, D.H., 1999,
preprint [astro-ph/9911343]
Geller, M.J., & Huchra, J.P., 1983, ApJS, 52, 61
Gioia I.M., et al., 1990, ApJS, 72, 567
Gioia I.M., et al., 2001, ApJL, 553, L105
Girardi, M., Borgani, S., Giuricin, G., Mardirossian, F., & Mezzetti,
Girardi M., Giuricin G., Mardirossian F., Mezzetti M., Boschin W.,
M., 2000, ApJ, 530, 62
1998, ApJ, 505, 74
Gladders M.D., Yee H.K.C., 2000, AJ, 120, 2148
Gonzalez A..H., Zaritsky D., Dalcanton J.J., Nelson A., 2001, ApJS,
137, 117
Goto, T., Sekiguchi, M., Nichol, R.C., Bahcall, N.A., Kim, R.S.J.,
Annis, J., Ivezic, Z., Brinkmann, J., Hennessy, G.S., Szokoly, G.P.,
& Tucker, D.L., 2001, submitted to AJ, astro-ph/0112482
Giuricin, G., Marinoni, C., Ceriani, L., & Pisani, A., 2000, ApJ, 543,
178
Haiman Z., Mohr J., Holder G., 2001, ApJ, 553, 545
Hattori M., Kneib J.P., Makino N., Prog. Th. Phys., in press [astro-
ph/9905009]
Helsdon, S.F., & Ponman, T.J., 2000, MNRAS, 319, 933
Henry J.P., Arnaud K.A., 1991, ApJ, 372, 410
Henry J.P., 2000, ApJ, 534, 565
Hernquist, L. & Katz, N., 1989, ApJS, 70, 419
Horner D.J., Mushotsky R.F., Scharf C.A., 1999, ApJ, 520, 78
Huchra, J.P., & Geller, M.J., 1983, ApJ, 265, 356
Huterer D., White M., preprint [astro-ph/0206292]
Ikebe, Y., Reiprich, T.H., Bohringer, H., Tanaka, Y., & Kitayama,
T., 2002, A&A, 383, 773
Jarrett, T.H., Chester, T., Cutri, R., Schneider, S., Skrutskie, M., &
Huchra, J.P., 2000, AJ, 119, 2498
Jenkins A., Frenk C.S., White S.D.M., Colberg J.M., Cole S., Evrard
A.E., Yoshida N., 2001, MNRAS, 321, 372
Jing Y.P., Mo H.J., Borner G., 1998, ApJ, 494, 1
Jones L.R., et al., 1998, ApJ, 495, 100
258, 1
Mahdavi, A., Bohringer, H., Geller, M.J., & Ramella, M., 2000, ApJ,
534, 114
Mahdavi, A., & Geller, M., 2001, ApJL, 554, L129
Marinoni, C. & Hudson, M.J. 2001, astro-ph/0109134
Marinoni, C., 2001, PhD thesis, University of Trieste
Marinoni, C., Hudson, M.J., & Giuricin, G., 2001, astro-ph/0109132
Markevitch M., 1998, ApJ, 504, 27
McLeod, B.A., & Rieke, M.J., 1995, ApJ, 454, 611
Moller, O., Natarajan, P., Kneib, J.P., Blain, A., 2001, ApJ,
submitted, [astro-ph/0110435]
Moore B., Frenk C.S., White S.D.M., 1993, MNRAS, 261, 827
Muanwong O., Thomas P.A., Kay S.T., Pearce F.R., 2002, preprint
[astro-ph/0205137]
Mulchaey, J.S., & Zabludoff, A.I., 1998, ApJ, 596, 73
Navarro J., Frenk C.S., White S.D.M., 1996, ApJ, 462, 563
Nevalainen J., Markevitch M., Forman W., 2000, ApJ, 532, 694
Nichol R., et al., 2000, in proceedings of MPA/MPE/ESO conference
"Mining the Sky", July 31 - August 4, 2000, Garching, Germany
[astro-ph/0011557]
Nolthenius, R.A., & White, S.D.M., 1987, MNRAS, 235, 505
Ostrander E.J., Nichol R.C., Ratnatunga K.U., Griffiths R.E., 1998,
AJ, 116, 2644
Ostriker J., Steinhart P.J., 1995, Nature, 377, 600
Peacock J.A., 2000, preprint [astro-ph/0002013]
Peacock J.A., Dodds S.J., 1996, MNRAS, 280, L19
Peacock J.A., Smith R.E., 2000, preprint [astro-ph/0005010]
Pearce, F.R. et al., 1999, ApJ, 521, 99
Pierpaoli E., Scott D., White M., 2001, MNRAS, 325, 77
Ponman T.J., Bourner P.D.J., Ebeling H., Bohringer H., 1996,
MNRAS, 283, 690
Postman M., et al., 1996, AJ, 111, 615
Postman, M., Lauer, T.R., Oegerle, W., Donahue, M., 2002, ApJ, in
press [astro-ph/0205513]
Press W.H., Schechter P., 1974, ApJ, 187, 452
Ramella, M., Diaferio, A., Geller, M.J., Huchra, J.P., 1994, AJ, 107,
1623
123, 2976
Ramella, M., Pisani, A., Geller, M.J., 1997, AJ, 113, 483
Ramella, M., Geller, M.J., Pisani, A., & da Costa, L.N., 2002, AJ,
Reblinsky K., Bartelmann M., 1999, a, 345, 1
Reiprich, T.H., & Bohringer, H., 2002, ApJ, 567, 716
Rines, K., Geller, M.J., Kurtz, M.J., Diaferio, A., Jarrett, T.H., &
Huchra, J.P., 2001, ApJ, 561, L41
Romer A.K., et al., 2000, ApJS, 126, 209
Rosatti P., Della Ceca R., Burg R., Norman C., Giacconi R., 1995,
ApJ, 445, L11
Scharf C.A., et al., 2000, ApJ, 528, L73
Schlegel, D.J., Finkbeiner, D.P., & Davis, M., 1998, ApJ, 500, 525
Schneider P., Kneib J-P., 1998, in ESA conference Proceedings of the
"Workshop on the Next Generation of Space Telescope: Science
Drivers & Technical Challenges" Liege, Belgium June 15-18, 1998
[astro-ph/9807091]
Scoccimarro R., Sheth R.K., Hui L., Jain, B., 2001, ApJ, 546, 20
Scodeggio M., et al., 1999, A&AS, 137, 83
Seljak U., 2000, preprint [astro-ph/0001493]
Shectman S., 1985, ApJS, 57, 77
Sheth R., Diaferio A., 2000, preprint
Skrutskie, M.F., Schneider, S.E., Stiening, R., Strom, S.E.,
Weinberg, M.D., Beichman, C., Chester, T, et al., 1997, in The
Impact of Large Scale Near-IR Sky Surveys, F. Garzon et al., eds.,
(Dordrecht: Kluwer) 187
Tully, R.B., 1987, ApJ, 321, 280
van Haarlem M.P., Frenk C.S., White S.D.M., 1997, MNRAS, 287,
817
Vikhlinin A., et al., 1998, ApJ, 503, 77
Voges, W., Aschenbach, B., Boller, Th., et al., 1999, A&A, 349, 389
Voit, G.M., Bryan, G.L., Balogh, M.L., Bower, R.G., 2002, ApJ in
press [astro-ph/0205240]
22
White, D.A., 2000, MNRAS, 312, 663
White M., 2001, A&A, 367, 27 [astro-ph/0011495]
White M., Hernquist L., Springel V., 2001, ApJ, 550, L129
White D.A., Jones C., Forman W., 1997, MNRAS, 292, 419
White M., Kochanek C.S., 2002, ApJ, in press [astro-ph/0110307]
White M., van Waerbeke L., Mackey J., 2002, ApJ, in press [astro-
ph/0111490]
White M., 2002, ApJS, in press [astro-ph/0207185]
White R., et al., 1999, AJ, 118, 2014
Wittman D., et al., 2001, ApJ,557, L89 [astro-ph/0104094]
Worthey, G., 1994, ApJS, 95, 107
Wu X-P., Chiueh T., Fang L-Z., Xue Y-J., 1998, MNRAS, 301, 861
Wu X-P., Fang L-Z., Xu W., 1998, A&A, 338, 813
Wu X-P., Xue Y-J., Fang L-Z., 1999, ApJ, 524, 22
Xue, J.-J., & Wu, X.-P., 2000, ApJ, 538, 65
Xu H., Jin G., Wu X.-P., 2001, ApJ, 553, 78
Yang X., Mo H.-J., van den Bosch F., 2002, preprint [astro-
ph/0207019]
Zaritsky, D., Nelson, A.E., Dalcanton, J.J., & Gonzalez, A.H., 1997,
ApJL, 480, 91
|
astro-ph/9812261 | 1 | 9812 | 1998-12-14T11:46:59 | Relativistic Distortions in the X-ray Spectrum of Cyg X-1 | [
"astro-ph"
] | We present the first significant detection of relativistic smearing of the X-ray reflection spectrum from the putative accretion disk in the low/hard state of Cyg X-1. The ionization state, and amount of relativistic smearing are simultaneously constrained by the X-ray spectra, and we conclude that the disk is not strongly ionized, does not generally extend down to the last stable orbit at 3 Schwarzschild radii and covers rather less than half the sky as seen from the X-ray source. These results are consistent with a geometry where the optically thick disk truncates at a few tens of Schwarzschild radii, with the inner region occupied by the X-ray hot, optically thin(ish) plasma. Such a geometry is also inferred from previous studies of the reflected spectrum in Galactic Black Hole transient sources, and from detailed considerations of the overall continuum spectral shape, suggesting that this is a robust feature of low/hard state accretion onto Galactic Black Holes. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 8 March 2018
(MN LATEX style file v1.4)
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
C. Done1, P.T. Zycki1,2
1 Department of Physics, University of Durham, South Road, Durham, DH1 3LE
2 Nicolaus Copernicus Astronomical Center, Bartycka 18, 00-716 Warsaw
8 March 2018
ABSTRACT
We present the first significant detection of relativistic smearing of the X -- ray
reflection spectrum from the putative accretion disk in the low/hard state of Cyg
X -- 1. The ionization state, and amount of relativistic smearing are simultaneously
constrained by the X -- ray spectra, and we conclude that the disk is not strongly ionized,
does not generally extend down to the last stable orbit at 3 Schwarzschild radii and
covers rather less than half the sky as seen from the X -- ray source. These results are
consistent with a geometry where the optically thick disk truncates at a few tens
of Schwarzschild radii, with the inner region occupied by the X -- ray hot, optically
thin(ish) plasma. Such a geometry is also inferred from previous studies of the reflected
spectrum in Galactic Black Hole transient sources, and from detailed considerations
of the overall continuum spectral shape, suggesting that this is a robust feature of
low/hard state accretion onto Galactic Black Holes.
Key words: accretion, accretion discs -- binaries: close -- black hole physics -- stars:
individual (Cygnus X-1) -- X -- ray: general -- X -- ray: stars
1
INTRODUCTION
Some of the strongest evidence for the existence of black
holes has come from recent ASCA (0.6 -- 10 keV) X -- ray ob-
servations of the shape of the iron Kα line in Active Galactic
Nuclei (AGN). AGN typically produce copious hard X -- ray
emission, and the iron line is formed from fluorescence as
these X -- rays illuminate the infalling material. The combi-
nation of Doppler effects from the high orbital velocities and
strong gravity in the vicinity of a black hole gives the line a
characteristically skewed, broad profile (Fabian et al., 1989).
This has been unambiguously identified from ASCA 0.6 -- 10
keV spectra of the AGN MCG -- 6 -- 30 -- 15, where the breadth
of the line profile implies that the accretion disk extends
down to at least 6Rg, the last stable orbit in a Schwarzschild
metric (Tanaka et al 1995; Iwasawa et al 1996).
As well as producing the line, some fraction of the illu-
minating hard X -- rays are reflected from the accretion flow,
producing a characteristic continuum spectrum. The am-
plitude of the line and reflected continuum depend on the
amount of material being illuminated by the hard X -- rays, its
inclination, elemental abundances and ionization state (e.g.
Lightman & White 1988; George & Fabian 1991; Matt, Per-
ola & Piro 1991; Ross & Fabian 1993; Zycki & Czerny 1994).
The mean amount of reflection and line seen in Seyfert 1
AGN is consistent with a power law X -- ray spectrum illumi-
nating an optically thick, (nearly) neutral disk, which sub-
tends a solid angle of ∼ 2π (Pounds et al 1990; Nandra
c(cid:13) 0000 RAS
& Pounds 1994), although it is now recognised that indi-
vidual objects show significant dispersion about this mean
(Zdziarski, Lubi´nski & Smith 1999).
This contrasts with the situation in the Galactic Black
Hole Candidates (GBHC). These are also thought to be pow-
ered by accretion via a disk onto a black hole, and, in their
low/hard state, show spectra which are rather similar to
those from AGN. However, the amount of reflection and iron
line is much less than would be expected from an accretion
disk which subtends a solid angle of 2π (e.g. Gierli´nski et
al 1997a), and the detected line is narrow, with no obvious
broad component (Marshall et al 1993, Ebisawa et al 1996,
hereafter E96). There are (at least) two possible explana-
tions for this: firstly that it is an artifact of the difference
in ionization state of the disk between GBHC and AGN, or
secondly that there is a real difference in geometry between
the supermassive accreting black holes and the stellar mass
ones.
Ionization differences seems at first sight to be an ex-
tremely attractive option. For accretion at the same fraction
of Eddington, the disk temperature should scale as M −1/4.
The GBHC inner disk is then expected to be a factor of
∼ 30 hotter than in AGN, and the higher expected ioniza-
tion state from the thermal ion populations gives a reflected
continuum and associated iron line that can be very differ-
ent to that from a neutral disk (Lightman & White 1988;
Ross & Fabian 1993; Matt Fabian & Ross 1993ab; Zycki et
al 1994; Matt Fabian & Ross 1996; Ross, Fabian & Brandt
2
C. Done & P.T. Zycki
1996). As well as increasing the energy of the iron edge (and
line), ionization also reduces the low energy opacity of the
disk, and so enhances its reflectivity. However, the iron edge
(whatever its energy) has a fairly constant cross-section, so
the net result is to increase the relative strength of the iron
edge feature in the reflected continuum. There is also the
possibility of the line being suppressed through resonant ab-
sorption over a (rather small) range in ionization states, so
the net result can be a prominent edge without obvious line
emission accompanying it. Such models were proposed as the
solution to the weak line emission seen in GBHC compared
to AGN (Ross & Fabian 1993; Matt et al., 1993a; Ross et al
1996).
However, detailed spectral fitting with photo -- ionized
reflected continuua shows that the ionization state of the
reflector is in general too low for the resonant absorption
process to be important. This argues against such models,
although relativistic shifts could affect the derived ionization
states (Matt et al 1993a, Ross et al., 1996). Another way to
suppress both line and reflected continuum is from a radial
dependance of ionization state. If the inner region of the
disk were so highly ionized that they produce no atomic
features then their contribution to the reflected continuum
is unobservable in the 2 -- 20 keV range. Here we fit data
from Cyg X -- 1 with a reflection model which includes both
ionization (as a function of radius) and relativistic smearing
in order to test whether the spectra are indeed affected by
resonant absorption and whether radial ionization structure
is important in masking the reflection signature.
We significantly detect relativistic smearing of the re-
processed features in Cyg X -- 1, and find that the disk is not
highly ionized. This apparently rules out both featureless
(extreme ionization) reflection and resonant absorption. The
line is not suppressed relative to the reflected continuum, it
is merely broadened by the relativistic effects which made
it difficult to detect in ASCA data (E96), where the ob-
served weak and narrow line emission probably comes from
the companion star. The derived covering fraction of the
relativistic reflector is substantially less than expected from
a flat disk. The most plausible explanation for this is that
the optically thin(ish) X -- ray corona fills a central 'hole',
so that less than half of the X -- ray flux intercepts the disk.
The amount of relativistic smearing generally is inconsistent
with the disk extending down to the last stable orbit at 6Rg.
These results from the shape of the reprocessed spectrum are
remarkably similar to the geometry derived from energetic
arguments from the continuum spectral shape (Gierli´nski et
al. 1997a; Dove et al 1997; Poutanen, Krolik & Ryde 1997).
Equally, they are very similar to the derived reflection ge-
ometry from spectra of transient GBHC during their decline
( Zycki, Done & Smith 1997; Zycki, Done & Smith 1998ab),
pointing to a single, fairly stable geometry for the low/hard
state GBHC.
2 THE SPECTRAL MODEL
The reflection model is described in detail in Zycki et al
(1998b). Basically it consists of the angle dependant reflec-
tion code of Magdziarz & Zdziarski (1995), with ion popu-
lations calculated as in Done et al (1992), as implemented
in the 'pexriv' model in XSPEC. This determines the ion
populations by balancing photoelectric absorption against
radiative recombination. The photo -- electric absorption edge
energies for iron were corrected from the rather approximate
values given by Reilman & Manson (1978) to those of Kaas-
tra & Mewe (1994). As noted by E96, these give a significant
difference in the derived ionization state. Photo -- ionization
rates are strongly dependent on the parameter ξ = L/(nr2)
ergs s−1 cm−1 (where L is the ionizing luminosity, here taken
to be between 5eV and 20 keV, n is the density and r the
distance of the material from the ionizing flux) and weakly
on spectral shape (assumed here to be a power law), while
recombination depends on temperature (assumed fixed at
106 K). The self -- consistent iron line is then calculated from
the Monte -- Carlo code of Zycki & Czerny (1994) and added
to the continuum. We assume fixed 'solar' abundances from
Morrison & McCammon (1983), except for iron, which is
free to vary from its tabulated value of Fe/H = 3.3 × 10−5.
Ionization as a function of radius can easily be treated
by adding together many radial rings with ionization ξ(r) ∝
rβ, where the relative covering fraction for each ring can be
calculated from the X -- ray irradiation Firr(r) ∝ r−3.
The relativistic smearing is taken from the XSPEC
'diskline' model, with slight modifications to take light
bending into account (Fabian et al 1989). This is param-
eterised by the inner and outer radius of the disk, Rin and
Rout, its inclination and (again) the irradiation emissivity,
Firr(r) ∝ r−3. We fix Rout = 104Rg (where Rg = GM/c2)
and fit for Rin, convolving the relativistic transfer function
with the total (line plus continuum) reprocessed spectrum.
For the multiple ionization state disk the relativistic transfer
function is calculated for each annulus, convolved with the
reprocessed spectrum from that ring and then co -- added to
get the total spectrum.
We assume a power law illuminating spectrum. This is
an approximation to the true shape of a Comptonised spec-
trum and it becomes a poor description at energies close to
the seed photon energy and close to the electron tempera-
ture. Typically in low/hard state of Cyg X -- 1 the electron
temperature is ∼ 100 keV (e.g. Gierli´nski et al 1997a), so
this will not give significant distortions to the spectral range
considered here, which is always ≤ 20 keV.
The situation with regards the soft photons is less
clearcut. The strongest component of the soft X -- ray contin-
uum can be modelled as a blackbody or disk blackbody with
a temperature of 0.1 − 0.2 keV (e.g. Ba luci´nska & Hasinger
1991; E96; di Salvo et al 1998). Again, this is far enough
away from the energy range considered in this paper (≥ 2
keV) for the distortions of the true Compton scattered spec-
trum from a power law to be negligible. However, there is
also a second soft component which extends to higher en-
ergies (E96, di Salvo et al 1998). E96 modelled this in the
ASCA data using a softer power law below a break energy
around 3 keV, but such a broken power law is rather un-
physical. Perhaps a more realistic model is one where soft
photons from the accretion disk (at ∼ 0.1 keV) are Comp-
tonised by a rather cool population of electrons with kT ∼ 1
keV as well as by the hot electrons with kT ∼ 150 keV. The
disk photons at ∼ 0.1 keV are energetically more important,
so it seems probable that this component at ∼ 1 keV does
not contribute significantly to the seed photons.
The physical origin of the second soft excess is entirely
unknown, and so is its geometry with respect to the disk and
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
3
spectral features from iron (see Figure 1). Including a neu-
tral reprocessed component inclined at 30◦ (Gies & Bolton
1986) gives a dramatic decrease in χ2
ν to 677/620, with a
further significant improvement in the fit if this is allowed
to be ionized (χ2
ν = 611/615). Allowing the iron abundance
to be a free parameter makes no significant difference, with
χ2
ν = 610/614 for AFe = 0.9 ± 0.2× solar.
Allowing the reprocessed spectrum to be relativistically
smeared gives a significantly better fit, with χ2
ν = 535/609.
We obtain constraints on the inner radius of the disk in
each spectrum, and these are somewhat larger than the 6Rg
expected for a disk extending down to the last stable or-
bit, assuming an illumination which is ∝ r−3. The derived
iron abundance is AFe = 1.8+0.7
−0.4× solar i.e. supersolar abun-
dances are preferred. This is consistent with the ∼ 2× solar
abundance of iron (given as Fe/H= 6 ± 1 × 10−5) inferred
for the stellar wind of the companion star from the strong
edge feature seen in the absorbed 'dip' spectra (Kitamoto
et al., 1984). The relativistic smearing allows more iron line
to be present, since it is broad rather than narrow, so it is
much less observable.
One potential problem may arise if there is a substan-
tial narrow line component from X -- ray illumination of flared
outer regions of the disk and/or the companion star (Basko
1978) and stellar wind. This could distort the relativistic line
fit, by adding a sharp core to the line profile. Such a line is
indeed detected in high resolution BBXRT and ASCA data
(Marshall et al 1993; E96). We allow for this by including a
narrow Gaussian line at 6.4 keV, whose normalisation is a
free parameter for each spectrum. This narrow component is
not significantly detected in the GSPC data, since it reduces
χ2 by only ∼ 1 for 5 additional parameters. The upper limits
on the derived line equivalent widths are typically less than
∼ 30 eV, easily compatible with the ASCA and BBXRT
detections. Clearly there should also be a reflection contin-
uum accompanying this line unless a substantial portion of
it arises from an optically thin stellar wind. We model this
by including an additional, unsmeared, neutral reprocessed
spectrum, assuming a fixed inclination of 60◦. Again, this
does not improve the fit (χ2
ν = 535/604), but we include
it for completeness. Details of this fit are given in Table 1,
where the iron abundance is frozen at its best fit value of
1.8× solar so that the error bars can be calculated for each
spectrum separately.
Several points are immediately apparent from this se-
ries of models. Firstly, the line strength is perfectly con-
sistent with solar, or even supersolar abundances. It is not
anomalously weak compared to the reflected continuum as
proposed by Ross et al (1996). The reflector is significantly
ionized, but not to the strength that would be required for
Auger ionization to operate (ξ ∼ 300 − 1000 for a power
law of Γ = 1.7 and temperature of 106 K). The derived
ionization state assuming an inclination of 30◦ is typically
that of Fe VIII-XII, incompatible with the Auger range of
Fe XVII -- XXIII. These ionization states are obtained de-
spite allowing for the effects of relativistic smearing. There is
no conspiracy, whereby the high Auger ionization is masked
by Doppler smearing and gravitational redshifts (Ross et al
1996). Instead the line and reflection continuum are per-
fectly consistent with each other, and with being produced
in a disk which is weakly ionized.
Figure 1. The ratio of each EXOSAT GSPC spectra of Cyg
X -- 1 to a best fitting power law model. Both excess counts at
the line energy and a deficit at the edge are seen, as in AGN.
The dotted line shows the same ratio for the GINGA -- 12 AGN
spectrum (Pounds et al 1990) for comparison.
hot plasma. It may or may not then produce its own reflected
component. We assume it does not, but note that even if it
does then its contribution to the crucial iron line and edge
region will be very small due to the low temperature of this
component.
In the following fitting we generally neglect data below
4 keV so as to avoid the complexities of the soft excess, and
fix the galactic column at 6 × 1021 cm−2.
3 CONSTANT IONIZATION REFLECTION
3.1 EXOSAT GSPC data
The EXOSAT GSPC data from Cyg X -- 1 give some of the
best spectra to date from this object, with the broad 2 -- 20
keV energy range of GINGA data but resolution comparable
to that of the ASCA GIS. We use the 5 GSPC spectra of
Done et al., (1992). The residuals to the best fit simple power
law for each GSPC spectrum are shown in Figure 1, together
with the residuals to a simple power law fit to the GINGA --
12 AGN spectrum of Pounds et al (1990). Contrary to claims
by Ross et al (1996), Cyg X -- 1 is not dominated by the edge
structure, but also shows a strong excess at the iron line
energy. Also, both line and edge appear equally diminished
in Cyg X -- 1 compared to AGN, although it is somewhat
misleading to compare these residuals directly since Cyg X --
1 is strongly absorbed by the interstellar medium and can
have a substantial contribution at 2 keV from the observed
soft excess.
We fit the 5 GSPC spectra simultaneously in XSPEC,
although only the reflection parameters of inclination and
iron abundance are constrained to be equal across all the
datasets.
We first fit the 4 -- 20 keV data with a power law. This
ν = 1309/625) due to the presence of
gives a poor fit (χ2
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4
C. Done & P.T. Zycki
Table 1. EXOSAT GSPC, GINGA and ASCA data from Cyg X -- 1. Error bars are ∆χ2 = 2.7, Galactic column is fixed at 6 × 1021 cm−2,
inclination at 30◦ and iron abundance at 1.8× solar
dataset
PL Γ
PL Norm
Ω/2πrel
GSPC 08
GSPC 09
GSPC 10
GSPC 13
GSPC 14
GINGA 91-1
GINGA 91-2
GIS 1
GIS 3
GIS 4
SIS 5
GIS 6
GIS 7
GIS 8
GIS 9
GIS 10
1.75+0.02
−0.01
1.56 ± 0.02
1.58 ± 0.02
1.75+0.03
−0.02
1.71 ± 0.02
1.60+0.05
−0.07
1.61+0.03
−0.08
1.87 ± 0.02
1.84+0.02
−0.01
1.71 ± 0.02
1.68 ± 0.03
1.68 ± 0.02
1.44+0.03
−0.02
1.66 ± 0.02
1.59 ± 0.02
1.60+0.03
−0.01
1.60
1.24
1.16
2.35
1.75
1.61
1.85
1.79
1.80
1.24
1.22
1.89
0.93
1.44
1.43
1.62
0.21+0.04
−0.05
0.19+0.04
−0.07
0.11+0.04
−0.06
0.24 ± 0.05
0.22 ± 0.04
0.23+0.09
−0.11
0.29+0.06
−0.17
0.05+0.03
−0.01
0.06+0.03
−0.02
0.1 ± 0.03
0.09+0.78
−0.04
0.02+0.09
−0.01
0.05+0.08
−0.02
0.06+0.06
−0.04
0.09+0.09
−0.08
0.06+0.11
−0.04
ξ
17+40
−16
24+90
−20
30+40
−26
31+60
−29
46+20
−42
60+640
−58
28+800
−28
2300+6000
−1600
3 ± 2.4 × 104
50+100
−28
3+32
−2.2 × 104
−0.01 × 103
−7.9 × 103
115+1600
6.70+193
8.4+42
−90
0+40
110+2000
−102
Rin(Rg)
Ω/2πnon−rel
χ2
ν
17+14
−8
13+9
7
1000−952
8+12
−2
6.0+6
10+∞
−4
28+∞
−22
16+7
−5
58+110
−31
90+120
−40
23+38
−11
31+210
−25
14+14
−6
37+40
−21
8+36
−2
64+100
−30
0+0.07
0+0.10
0+0.12
0+0.11
0+0.03
a
0.07+0.03
−0.07
0+0.1a
0+0.03
0.04 ± 0.04
0+0.05
0.05+0.08
−0.04
0.13+0.06
−0.05
0+0.03
0.05 ± 0.05
0.13+0.05
−0.06
0.08 ± 0.06
89.2/118
92.9/98
69.3/104
174.8/176
108.7/109
10.2/21
9.8/21
107.8/79
92.0/79
142.8/79
51.8/70
84.4/79
54.3/79
108.2/79
68.1/79
95.8/79
a Parameter restricted to the range 0-0.1 since it is not well constrained
3.2 GINGA data
3.3 ASCA data
The GINGA data give consistent results, though at much
lower significance. There were 3 epochs in which Cyg X --
1 was observed, the first of which (August 1987) was mis-
pointed and the remaining two (in May 1990 and June 1991)
being in an unusual mode designed to increase the spectral
range out to high energies. In this mode the spectral resolu-
tion is reduced by about a factor 2 since the same number
of channels cover twice the bandpass. The resolution is then
more like 30 per cent at iron, rather than 17 per cent in the
usual mode. This is a crucial difference, since the broadest
features expected from the inner disk are of order 30 per
cent if the disk extends down to the last stable orbit, so
these data are unlikely to be able to constrain the amount
of relativistic smearing. Ironically, the larger bandpass is not
actually of use either, as data beyond 30 keV are limited by
uncertainties in background subtraction.
We use datasets 1 and 2 (the unabsorbed spectra)
from the June 1991 observations shown by Gierli´nski et al
(1997a). These are fit in the 4 -- 30 keV range with the same
model as above. The poor resolution of the data means that
the relativistic smearing is not significantly detected, so the
relativistic and non -- relativistic reflected spectra are degen-
erate. From the EXOSAT results and from theoretical ex-
pectations of the amount of reflection from the companion
star (Basko 1978) we limit the solid angle Ω/2πnon−rel ≤ 0.1.
With this constraint we obtain the results shown in Table 1.
Again the data have a line which is easily consistent with the
amount of reflection seen. The observed ionization parame-
ter of the reflecting material is generally low (though poorly
constrained) even when relativistic distortions are included.
The ASCA SIS and GIS data used here are those published
by E96, except that we neglect the October 1993 SIS spectra
(spectrum 2 in E96) where there may be a residual problem
with the resolution (E96). Again we use only data above 4
keV so as to avoid the complexities of the soft X -- ray excess,
and so that the fits can be compared with the results of E96.
These data clearly show a narrow line component as well as
a complex edge structure which has been modelled by an
ionized reflection (E96). However, their reflected continuum
model did not include the self consistent iron line emission
nor relativistic smearing.
The SIS spectrum has the best energy resolution
(November 1993, denoted spectrum 5 in E96) so we first ex-
amine the line profile in these data. We fit this with the self --
consistent iron line and continuum reflection code, together
with a power law continuum and galactic NH = 6 × 1021,
firstly assuming an inclination of 30◦. Without relativistic
effects the fit is much worse than that shown by E96, with
χ2
ν = 98/71, showing that the narrow line cannot be the
corresponding fluorescence from the material giving rise to
the observed reflection continuum, irrespective of the iron
abundance (best fit value of unity) and ionization (ξ ∼ 5).
Including relativistic effects gives a much better fit, with
χ2
ν = 56/70, comparable with that of E96. Including a sepa-
rate narrow line gives an even better fit, with line energy and
equivalent width of 6.46 keV and 10 eV (χ2
ν = 48/68). How-
ever, physically we expect a self -- consistent reflected contin-
uum to accompany this line. Replacing the line with the
total reprocessed spectrum from neutral material where rel-
ativistic effects are not important gives an equally good fit
with χ2
ν = 51/69. This fit (shown in Figure 2) is very dif-
ferent to the one derived from the EXOSAT data. The ion-
ization is so high (ξ ∼ 4 × 104) that the 'edge' feature seen
in the data is actually modelled by the drop following the
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
5
Figure 2. The best fit to the ASCA SIS 5 spectrum assuming
an inclination of 30◦, with iron abundance free to vary between
1 − 2× solar. The relativistic reflector has ξ = 3.0+37
−2.2 × 104,
Rin = 23+40
−0.04, while the neutral unsmeared
reflector has Ω/2π = 0.05+0.10
−0.04, for an illuminating power law
with α = 0.68 ± 0.03 (χ2
−11 and Ω/2π = 0.09+0.81
ν = 51.8/70).
relativistically smeared (Rin ∼ 22Rg) ionized line, and the
real edge is shifted outside of the observed energy range.
The GIS spectra give similar results. We fit these using
the most recent GIS response matrices (as of March 1995,
gis2v4-0.rmf) as opposed to the earlier response used by
E96, and results are detailed in Table 1. Again, the derived
ionization paramter is generaly very high -- only the GIS 4
and GIS 9 spectra give results that are anything like the low
ionization EXOSAT and GINGA solutions.
3.4 Inclination
This general mismatch between the ionization of the repro-
cessor derived from ASCA and EXOSAT/GINGA data is
suggestive of a systematic problem in the spectral model
fitting, most likely due to the difference in bandpass. In EX-
OSAT and GINGA the high energy continuum shape of the
reprocessed spectrum helps constrain its ionization, whereas
in ASCA only the iron features can be used. The detailed
shape of the line and edge is a strong function of inclination
as well as of ionization. At high inclinations Doppler shifts
prevail over gravitational and transverse redshift, shifting
the line and edge to higher energies as well as giving sub-
stantial broadening. This is to zeroth order the same effect as
ionization. However, the detailed shape of a low inclination,
ionized reflection spectrum is rather different to a less ion-
ized reflection spectrum at higher inclination, so these two
parameters can be disentangled given good enough data.
Figure 3a shows how χ2 varies for simultaneous spectral
fitting of the five EXOSAT datasets as the disk inclination
is changed from 30 − 66◦. Clearly there is a significant pref-
erence in the data for an inclination higher than 45◦ (cosine
smaller than 0.7). The 3σ (∆χ2 = 9) limit to the inclination
from this modelling is i ≥ 40◦. This is rather higher than
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 3. The change in χ2 (goodness of fit parameter) as a
function of inclination of the reflecting material. Panel (a) shows
results from a simultaneous fit of all the EXOSAT GSPC spec-
tra, with the solid line showing fits to the data from 4 -- 20 keV
while the dotted line uses the full 2 -- 20 keV bandpass including a
(diskblackbody) model for the soft excess. The latter have 72 sub-
tracted from them in order to appear on the same scale. Panel (b)
shows the ASCA GIS 3 (solid line) and ASCA SIS 5 (dotted line)
4 -- 10 keV spectral results, where the latter have been boosted by
20 in order to appear on the same scale. Clearly the data which
have significant bandpass beyond 9 keV favour inclinations higher
than 30◦.
the most probable inclination inferred from optical studies
of 28◦ − 38◦ (Gies & Bolton 1986), but within the limit set
by the most recent work of ≤ 55◦ (Sowers et al 1998). At
the best fit GSPC inclination of 53◦ the iron abundance is
2+0.6
−0.4× solar (χ2
ν = 522/604). We check that this is not an
artifact of ignoring the soft X -- ray excess by including data
down to 2 keV and approximating it with a diskblackbody
6
C. Done & P.T. Zycki
spectrum (XSPEC model diskbb). The dotted line in figure
3a shows that the data still significantly favour the higher
inclination solutions.
Despite their higher spectral resolution, the ASCA SIS
are completely unable to constrain the inclination (Figure
3b, dotted line). Figure 4a shows this degeneracy of SIS
solutions in more detail. At low inclinations the data select
uniquely the high ionization solution (shown in Figure 2),
while at high inclinations an equally good solution can be
found at both high and low ionization. Figures 5a and b
show these high inclination solutions -- the low ionization
one has the 'pseudo -- edge' in the data matched by the real
edge while at high ionization it is matched by the drop from
the broadened ionized line. Figures 5a and b also show the
extension of these models out to 20 keV. It is obvious that
these solutions can be distinguished with higher energy data.
The GIS data have just enough extra bandpass at high
energies (above 9 keV) to break this degeneracy. The dashed
line in figure 4b shows the change in χ2 with ionization for
cos i = 0.86 and 0.6 for the GIS spectrum 3 of E96. Not only
is the ionization uniquely determined for a given inclination,
the data also are able to distinguish between the high ion-
ization/low inclination and low ionization/high inclination
solutions, clearly preferring the latter. The correspondence
with the GSPC data is clear, so we fix the inclination of
the relativistic reflector at 53◦ (cosine of 0.6) and re -- fit all
the data, using the best fit GSPC iron abundance of twice
solar. These results are shown in Table 2. For the EXOSAT
GSPC (and GINGA) data the fits are only subtly different --
the derived ionization state is somewhat lower while the in-
ner radius is somewhat larger, both because of the stronger
doppler effects in the higher inclination models. However,
for the ASCA data the high inclination solutions are very
different from those in Table 1. Firstly, they are statistically
better -- the total χ2 from all the GIS data for the high incli-
nation solution is 698.5/632 as opposed to 753.4/632 for the
low inclination one. Secondly, the derived ionization drops
dramatically, and now corresponds to that seen in the EX-
OSAT GSPC data. All the data now give similar derived
parameters, irrespective of bandpass. Only the SIS 5 and
GIS 6 spectra allow a high ionization solution.
Unlike the variable gain EXOSAT GSPC data, the GIS
data are all taken in the same mode, so their spectral residu-
als can be co -- added in order to illustrate the features which
make one model better than another. The data are fit si-
multaneously, with only the reflection parameters of iron
abundance (constrained to be between 1 -- 2× solar) and in-
clination fixed between the datasets. The top panel in Figure
6 shows the results from fitting a reflector which is ionized
but not relativistically smeared for an inclination of 30◦.
There are clear broad residuals present around the iron line
energy. A much better fit is obtained if the 30◦ ionized re-
flector is allowed to be relativistically smeared (and a non --
smeared, unionized reflector is also added), as shown in the
middle panel. However, the bluest features of the residuals
are not able to be adequately fit in these models, despite the
rather high derived ionizations (see table 1), due to the pre-
dominant redshift at 30◦. Higher inclination solutions give
a stronger doppler shift to the line, so matching the data
rather better (bottom panel).
For the high inclination solutions, all the data show
that the covering fraction of the relativistic reflector is al-
Figure 4. The change in the goodness of fit parameter χ2 as a
function of ionization state of the reflector for inclinations of 30◦
and 53◦. Panel (a) shows the results for the SIS 5 spectrum. For
an inclination of 30◦ the data uniquely select the high ionization
solution while at 53◦ there are equally good solutions at both low
and high ionization. All these three solutions are of comparable
statistical likelihood. Panel (b) shows the contrasting situation
derived from the GIS 3 spectrum. Here the data uniquely select
the high inclination, low ionization solution.
ways much less than unity, the derived inner disk radius is
generally significantly larger than 6Rg, and its ionization
state is rather low. The solid angle subtended by the narrow
reprocessor is always consistent with being the Ω/2π ≤ 0.1
expected from the companion star. This should vary with or-
bital phase (Basko 1978), although the data here give error
bars on the derived parameters which are too large to con-
strain this. Orbital variability in the narrow reflected com-
ponent would provide an explanation for the observed 4%
peak to peak quasi -- sinusoidal variability seen in the high en-
ergy BATSE (20 -- 100 keV) lightcurve of Cyg X-1 (Paciesas
et al 1997). The reflected component should be small (basi-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
7
Figure 6. Co -- added χ2 residuals from fitting the 8 GIS spectra.
The top panel shows results from fitting non -- smeared reflection
to the data. There are clearly broad residuals left around the iron
line energy. The mid panel shows that these are much reduced
by allowing the reflector to be a relativistic disk (including non --
smeared, unionized reflection from the companion star). However,
for the assumed inclination of 30◦, the highest energy residuals
are not matched by the models. The lower panel shows that these
can be fit at higher inclinations, where doppler blueshifts can
become stronger than the prevailing redshifts. The feature at 4.7
keV in all the plots is a residual from the changing response of
the GIS detectors at the Xe L edge.
10), even allowing for twice solar iron abundances. The rea-
son for the difference is that the fit is somewhat sensitive
to changes in the model, and we note that our description
gives better χ2 (by up to 20) than the fits of E96.
4 IONIZATION AS A FUNCTION OF RADIUS
The data clearly prefer a solution in which the covering frac-
tion of the reflector is extremely low, typically only 20 -- 30
per cent of what would have been expected from a disk which
covered half the sky as seen from the X -- ray source, and the
derived inner radius is generally inconsistent with the last
stable orbit. Both these results can easily be explained in
a scenario where the optically thick disk terminates before
the last stable orbit, with the inner radii instead being filled
with X -- ray hot optically thin(ish) plasma. However, a series
of papers (e.g. Ross et al 1996) have explored the possibility
that high ionization of the reflector supresses the derived
amount of reflection. We have shown above that this does
not work for Cyg X -- 1 when the reflector is modelled by
−0.04, Rin = 14 ± 4 and Ω/2π = 0.45+0.80
Figure 5. Two equally acceptable decompositions of the SIS
5 spectrum of Cyg X -- 1, assuming an inclination of 53◦ with
iron abundance free to vary. Panel (a) shows the low ioniza-
tion reflected continuum, with a relativistic disk of ionization
ξ = 0.04+25
−0.10 (χ2
ν =
49.7/69), for an illuminating power law with α = 0.73+0.16
−0.04 and
neutral unsmeared reflection Ω/2π = 0.05+0.10
−0.03. The iron abun-
dance is completely unconstrained between 0.5 − 3× solar. Panel
(b) shows the high ionization solution, where the relativistic disk
has ξ = 3+34
−0.11, with
Ω/2π = 0.07 ± 0.06 of unsmeared neutral reflection and a power
law with α = 0.70 ± 0.04 (χ2
ν = 50.0/69).
−2.6 × 104, Rin = 55+60
−30 and Ω/2π = 0.18+4.0
cally zero) at phase 0 when the companion star is in front of
the black hole, and have its maximum of Ω/2π ∼ 0.1 viewed
at an inclination of 90 −i at phase 0.5, giving a 5% peak -- to --
peak variability in the 20 -- 100 keV flux. We do not require
an additional narrow reflected component from the outer
disk, contrary to the conclusions of E96. Their fits to these
data gave some spectra which required a much larger narrow
component than could easily be explained by the compan-
ion star alone (e.g. GIS data from November 1994, spectrum
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
8
C. Done & P.T. Zycki
Table 2. EXOSAT GSPC, GINGA and ASCA data from Cyg X -- 1. Error bars are ∆χ2 = 2.7, Galactic column is fixed at 6 × 1021 cm−2,
inclination is 53◦ and the iron abundance is 2× solar
dataset
PL Γ
PL Norm
Ω/2πrel
ξ
Rin
Ω/2πnon−rel
GSPC 08
GSPC 09
GSPC 10
GSPC 13
GSPC 14
1.77 ± 0.02
1.58+0.02
−0.01
1.58+0.02
−0.01
1.78+0.02
−0.03
1.74+0.01
−0.02
GINGA 91-1
GINGA 91-2
1.63 ± 0.03
1.63 ± 0.03
GIS 1
GIS 3
GIS 4
SIS 5
GIS 6
GIS 7
GIS 8
GIS 9
GIS 10
1.91 ± 0.02
1.91 ± 0.02
1.74 ± 0.02
1.74 ± 0.04
1.69+0.03
−0.02
1.50 ± 0.03
1.66+0.02
−0.01
1.58 ± 0.02
1.61 ± 0.02
1.65
1.28
1.16
2.46
1.85
1.60
1.92
1.89
2.08
1.28
1.46
1.94
1.01
1.45
1.42
1.63
0.34+0.05
−0.14
0.26+0.09
−0.12
0.13+0.05
−0.07
0.34+0.12
−0.13
0.39 ± 0.06
0.33+0.15
−0.09
0.44+0.10
−0.21
0.36+0.12
−0.13
0.53+0.09
−0.17
0.31 ± 0.08
0.54+0.17
−0.13
0.10+0.17
−0.09
0.48+0.07
−0.11
0.09+0.05
−0.07
0.04+0.13
−0.04
0.04+0.09
−0.02
0.2+17
−0.2
0+0.3
21+70
−20
0+3
0+0.4
4+30
−4
3+45
−3
−0.02
2.8+40
−2.7
0.02+17
0+1
0+20
30+1e5
−30
0+10
50+2000
−45
0+∞
700+1700
−680
30+18
−11
28+22
−15
1000−970
15+15
−9
14+7
−4
40+∞
−30
30+∞
−22
20+6
−5
9.9+6
−1.9
6.0+2
13+4
−3
18+30
−12
14+5
−4
60+170
−44
51+∞
−45
100+220
−66
0+0.14
0.05+0.11
−0.05
0+0.13
0.09+0.10
−0.09
0+0.05
a
0.1−0.1
0+0.1a
0.01+0.04
−0.01
0.08+0.03
−0.04
0.09 ± 0.03
0.05 ± 0.03
0.12 ± 0.03
0+0.02
0.05+0.05
−0.04
0.12+0.04
−0.12
0.09+0.04
−0.05
χ2
ν
86.1/118
96.3/98
69.7/104
172.6/176
98.0/109
12.1/21
8.8/21
93.5/79
69.8/79
134.4/79
49.9/70
81.4/79
45.7/79
108.4/79
71.7/79
93.6/79
a Parameter restricted to the range 0-0.1 since it is not well constrained
Table 3. GSPC 08, SIS 5 and GIS 3 spectra fit to the radial ionization model with ξ(r) ∝ r−1.5. Error bars are ∆χ2 = 2.7, Galactic
column is fixed at 6 × 1021 cm−2. Inclination is fixed at 30◦ and 53◦ for each spectrum, while the iron abundance is allowed to vary between
1 -- 2× solar (so there is an extra degree of freedom compared to table 2).
Spectrum
PL Γ
AFe
Cos(i)
Ω/2πrel
ξ(Rin)
Rin
Ω/2πnon−rel
GSPC 08
GSPC 08
GIS 3
GIS 3
SIS 5
SIS 5
SIS 5
1.77+0.01
−0.03
1.77 ± 0.04
1.91+0.09
−0.02
1.85+0.01
−0.02
1.75+0.13
−0.04
1.69+0.05
−0.03
1.67+0.05
−0.03
2.0
1.2
2
2
2
2
2
0.60
0.86
0.60
0.86
0.6
0.6
0.86
0.34+0.04
−0.14
0.22+0.04
−0.03
0.53+0.40
−0.11
0.06 ± 0.01
0.56+0.59
−0.12
1.08+∞
0.11+∞
−0.99
−0.06
3.0−2.1
−0.01
2+50
−2
65+160
−56
0.01+11
+3.5 × 104
0.1+30
−0.1
−1.37 × 106
−1.5 × 105
1.40+2.1
1.8+38
27+17
−8
20+13
−7
9.5+3
−2.5
34+27
−12
13+6
1734
−10
14+16
−6
−3.5
0.01+0.12
−0.01
0+0.09
0.09+0.07
−0.04
0+0.06
0.06+0.08
−0.03
0.11+∞
0.05+∞
−0.06
−0.04
χ2
ν
86.2/117
88.5/117
69.7/78
89.7/78
49.4/69
50.7/69
52.0/69
material with constant ionization parameter. However, per-
haps there is a radial range of ionization states. If there is
an extremely ionized inner disk then it produces no spectral
features, so neither its covering fraction nor its relativistic
smearing can be seen.
We test this idea by dividing the disk into 10 radial
zones, again with the illumination ∝ r−3, and the ionization
ξ(r) ∝ r−1.5 as appropriate for a gas pressure dominated
disk. Due to computational time we only demonstrate this
model on a single spectrum from each instrument: EXOSAT
GSPC 08, ASCA GIS 3 and ASCA SIS 5. The iron abun-
dance is a free parameter (within the range 1 − 2× solar),
and we include the possibility of a neutral reflected compo-
nent from the star/outer disk. The fits are repeated for fixed
inclination of cos i = 0.86 and 0.6.
Details are given in Table 3, and are not significantly
different from those derived from the single ionization model.
The inferred covering fraction for the disk is still low, and the
disk inner radius is inconsistent with the last stable orbit.
We explicitally searched for a high ionization solution for the
cos i = 0.6 fits, but these always gave a significantly worse
χ2 except for the SIS 5 data.
The reason for the failure of the multi -- zone ionization
model is that the mean ionization state observed is rather
low. Between this and the extreme ionization states required
to make the reflector invisible are the intermediate/high ion-
ization states. These give a strong and distinctive spectral
signature, especially those where iron has only one or two
electrons left (H and He -- like, FeXXVI and XXV) which pro-
duce intense line emission at 7.0 and 6.7 keV, and (more im-
portantly) a deep edge at 9.2 and 8.6 keV (Ross & Fabian
1993). Even allowing for strong distortion by relativistic
smearing (Matt et al., 1993a; 1996; Ross et al., 1996) these
features are not seen in the data. It is not then possible to
go smoothly from the observed rather low ionization states
to the extreme ionizations required to hide reflection. Only
if there is an abrupt ionization transition can this model be
sustained.
Again, we note that good data at energies ≥ 9 keV are
required in order to constrain the amount of highly ionized
reflecting material. Radial ionization models with high inner
disk ionization can be fit to the SIS data for any inclination,
but are again ruled out by the GIS 3 spectrum (see table 3).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
9
5 DISCUSSION
From the above fits it is clear that the relativistic smear-
ing of the reflected spectrum is significantly detected in Cyg
X -- 1, confirming the presence of an optically thick material
at small radii -- the putative accretion disk. The covering
fraction of the disk is substantially less than unity (see Ta-
bles 1-3), irrespective of inclination, ionization and radial
gradients in ionization.
The ionization of the reflector and its inclination to the
line of sight can be constrained with moderate spectral reso-
lution data and bandpass extending beyond 9 keV. The data
show strong preference for a low ionization reflector inclined
at rather more than 30◦ to the line of sight: the ASCA GIS
3 spectrum gives a range of 41◦ − 61◦, while the EXOSAT
GSPC data imply angles greater than ∼ 47◦ (both at the
90% confidence limit). These inclinations are determined by
the detailed shape of the iron line in the data: for low in-
clinations the theoretical line profile is strongly affected by
gravitational redshift so that higher ionization solutions are
needed to fit the data. But in higher ionization solutions
there is a large energy separation between the line and edge
(however smeared they become) which is not seen in the
data. The spectral features observed require that the line
and edge blend together, i.e. that the ionization is moder-
ate to low. The reflector then is required to be viewed at a
moderate inclination angle so that the doppler blueshift can
counteract the gravitational redshift. However, these inclina-
tion are rather larger than the often quoted probable range
of 28◦ − 38◦ (Gies & Bolton 1986) derived from optical data,
although within the firm upper limit to the inclination from
these studies of i ≤ 55◦ (Sowers et al 1998). The lack of a well
determined optical orbital solution is due to the complexity
of the accretion flow from the supergiant companion in this
system. The companion probably does not completely fill
its Roche lobe, but instead has a strong stellar wind which
is gravitationally focussed by the Roche potential (see e.g.
numerical simulations by Blondin, Stevens & Kallman 1991)
While the iron line is strongly distorted by the relativis-
tic effects, the extremely broad components from material
at 6Rg are not present at the level expected from a disk
extending down to the last stable orbit with line emissivity
∝ r−3. Such a model is only possible if the inner disk has
an abrupt increase in ionization state to the level at which it
becomes completely ionized so produces no atomic spectral
features. We stress that a continuous (power law) ionization
gradient as expected from the intense X -- ray illumination is
insufficient to do this, since between the observed rather low
ionization state of the reflector and the extreme ionizations
required to make the disk completely reflective there are the
high ionization states, where iron has only 1 or 2 electrons
left. These ions produce copious Kα line emission at 6.7 and
7.0 keV, and strong edges at 8.8 and 9.2 keV, which are not
seen in the data. Similarly, these high ionization states are
precisely the ones expected from collisional ionization from
the accretion disk temperature. The observed soft excess is
dominated by a component at 0.1 − 0.2 keV which cannot
completely ionize iron. Even the secondary soft emission at
kT ∼ 1 keV is far removed from the inner K shell energy
of 8.8 and 9.2 keV for He and H like iron. This then rules
out the idea of a simple correspondence between AGN and
GBHC where both have disks which extend down to 6Rg
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
but in the GBHC the disk is ionized simply because of the
higher accretion disk temperature.
The optically thick disk must do something at the de-
rived Rin > 6Rg (assuming that the line emissivity is ∝ r−3;
Zycki et al 1998b). Either the inner disk disappears com-
pletely or it jumps discontinuously to a completely ion-
ized state at this radius. Similar results are seen for the
low/hard state GBHC transient sources ( Zycki et al 1997;
1998ab), showing that this is a robust feature of low/hard
state GBHC spectra.
Figure 7 shows the plot of spectral index versus the de-
rived covering fraction, ionization and inner radius of the
reflector for the reflector models inclined at 53◦ (not includ-
ing the poorly constrained GINGA data). The covering frac-
tion is not consistent with a constant (χ2
ν = 60/14), showing
that the geometry must be changing. Apart from the hardest
spectral index data (GIS 7, which incidentally is the spec-
trum taken furthest off axis, and so has a large vignetting
correction, E96) the data are consistent with a trend in
which a hard spectral index is associated with smaller re-
flected fraction, similar to results from some of the X -- ray
transients ( Zycki et al 1997; 1998ab). This can be qualita-
tively explained by a geometry in which there is a variable
radius inner hole in the cool disk which is filled with X -- ray
hot plasma, although the error bars on the radius derived
from relativistic smearing are too large to be able to con-
strain these ideas. A disk which extends further down into
the gravitational potential could subtend a larger solid an-
gle to the source. The soft photon flux from the disk would
also increase, both because the disk luminosity is dependent
on its inner radius, and also from the larger fraction of hard
X -- rays which can be reprocesssed. This would give stronger
Compton cooling of the hot electrons and so a steeper X -- ray
spectrum.
The typical derived inner radius of the disk is Rin =
20Rg. The luminosity released by viscous dissipation from
20Rg to 6Rg is rather less than 0.75 that released from ∞ to
20Rg (see e.g. equation 5.19 in Frank, King & Raine 1992).
Simple energetics then give the expectation that the X -- ray
hot plasma should carry ∼ 0.75× the luminosity of the soft
photons from the disk. However, the opposite is observed:
the hard component (the X -- ray hot plasma) is over twice as
luminous as the soft component from the disk e.g. di Salvo
et al., (1998).
However, this geometrical interpretation is clearly not
unique, and many others have been proposed (see e.g.
the review by Poutanen 1998). If the hot X -- ray emitting
plasma forms a (plane parallel) corona above the disk (e.g.
Haardt & Maraschi 1993) then the reflected spectrum has to
cross the corona before escaping to the observer. A fraction
exp(−τ / cos i) (where τ is the optical depth and i the inclina-
tion of the disk -- corona to the observer) will interact again
with the hot electrons in the plasma. The reflected spec-
trum then forms another source of seed photons for Comp-
ton scattering, and the scattered fraction joins the Compton
continuum and so is unrecognisable as reflected flux. Such
models were shown to work for Cyg X -- 1 data up to 20 keV
(Haardt et al 1994), but they predict that there should be a
strong break in the spectrum beyond this energy. Seed pho-
tons from the disk are not isotropic, so the first few Compton
scattering retain the imprint of this anisotropy and give a
somewhat flatter spectrum which then joins onto the more
10
C. Done & P.T. Zycki
Figure 7. The relativistic reflection spectrum parameters as a
function of the intrinsic spectral index for the model with reflector
inclination of 53◦ (table 2). The top panel shows the amount of
reflection, while the second panel shows its ionization state and
the third shows its inner radius.
isotropic higher order scatterings (Haardt & Maraschi 1993).
This anisotropy break is not present in broad band spectra
from Cyg X -- 1, where the data extend past 20 keV, which
rules out such models (Gierli´nski et al 1997a).
The lack of an anisotropy break seems to be an in-
surmountable problem for any model which has the X -- ray
emission region above a disk, including the 'active regions'
models which may be more appropriate if the X -- ray regions
are energised by localised magnetic reconnection events (e.g.
Haardt, Maraschi & Ghisellini 1994). However, these models
are very attractive given that accretion disk viscosity is now
thought to be driven by magnetic reconnection. But even if
the anisotropy break can be masked (perhaps by the elec-
trons also being anisotropic or by a distribution of electron
temperatures) there are further problems for models where
the X -- ray region is above the disk. If edge effects are not im-
portant (approximately plane parallel geometry) then about
half of the hard X -- rays are intercepted by the disk. The ma-
jority (whatever is not reflected) of this flux is thermalised in
the disk and re -- emerges as soft photon seeds for the Comp-
ton scattering. These soft photons all have to pass through
the corona in a plane parallel geometry, resulting in strong
cooling. Without a strong anisotropy break then the pre-
dicted spectra are too soft to match the observed low/hard
spectra from Cyg X -- 1 (Dove et al 1997, Poutanen et al.,
1997). This forces us to the 'active regions' models, where
not all flux of soft photons from the disk will pass through
the corona as edge effects are important. But in these models
there is then no overlying corona that can reduce the amount
of reflection seen. If edge effects are important, then much
of the reflection is produced from regions of the disk that
are not underneath the corona. Reflection is not then su-
presssed by Compton scattering in the corona, and there is
the strong prediction that we should see Ω/2πrel = 1, which
is plainly not the case.
These are serious problems to be faced by all models
which have the X -- ray region above the disk. The continu-
ous disk -- corona models can reduce the amount of reflection
seen, but predict a large anisotropy break which is strongly
ruled out by the data. Even if there is some way to mask the
anisotropy break then these models produce spectra which
are too soft to match what is seen. The 'active regions' mod-
els can produce hard spectra, but cannot supress reflection
and again predict an anisotropy break which is not seen.
Compared to these drawbacks, the central source geometry
seems an attractive option.
6 EQUILIBRIUM ACCRETION FLOW
SOLUTIONS
There are several solutions of the equations for accretion of
material onto a compact object. The most well known of
these is that of Shakura & Sunyaev (1973; hereafter SS),
which is the geometrically thin, optically thick accretion
disk. In this solution, the gravitational energy released is
radiated away locally as (quasi)blackbody radiation. How-
ever, other solutions are also possible: the energy can instead
be radiated by Compton scattering in an optically thin(ish),
geometrically thick(ish) flow (Shapiro, Lightman & Eardley
1976). Alternatively, the energy released need not be effi-
ciently radiated. It can be carried along with the flow (ad-
vected) and swept into the black hole. These inefficient, ad-
vectively cooled flows can either be optically thin (ADAFs:
Narayan & Yi 1995; Abramowicz et al 1995), or optically
thick (Abramowicz et al 1988), depending on the mass ac-
cretion rate. In general the solutions coexist: at any radius
there are several different equilibrium flow configurations for
a given mass accretion rate and viscosity (see e.g. Chen et
al., 1995).
6.1 Shakura -- Sunyaev Accretion
Why should the inner disk radius change, and not extend
down to 6Rg ? The 'standard' accretion disk model devel-
oped by SS gives the flow structure in the limit when the
gravitational energy is released in optically thick material.
The cooling is very efficient (blackbody radiation), so all
the energy released locally can be radiated locally, giving a
cool, geometrically thin disk structure. This becomes unsta-
ble when the total pressure is dominated by radiation pres-
sure (Shakura & Sunyaev 1976), giving a natural inner disk
cutoff. For steady state models these predict that the inner
radius should move outwards as the luminosity increases (as
the radiation pressure dominated region increases: SS), in
conflict with the observed trend in the transient GBHC to
show higher reflected fractions and smaller inner disk radii
at higher luminosities ( Zycki et al 1997; 1998a,b).
Another, more general problem with the SS accretion
disk structures is that such models give temperatures of or-
der 1 keV for GBHC at high accretion rates but are quite
unable to explain the observed (hard or soft) power law
tail out to energies ≥ 100 keV. Either there are parts of
the disk flow in a different configuration to that of SS, or
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
11
there is some non -- disk structure such as a corona powered
by magnetic reconnection (e.g. Haardt, Maraschi & Ghis-
ellini 1994). However, this latter possibility (a cool disk un-
derneath a plane parallel or hemispherical X -- ray emitting
corona) seems to be ruled out by detailed spectral fitting
(see Section 6). Instead it seems that the spectrum implies
a geometrical constraint that the X -- ray emitting plasma and
disk occupy more or less separate regions, such as an outer
disk and inner (quasi -- spherical ?) X -- ray source (Gierli´nski
et al 1997a, Dove et al 1997, Poutanen et al., 1997).
6.2 Advective Accretion
Recently there has been much excitement about the possi-
bility that another stable solution of the accretion flow may
explain the hard X -- ray data. Below a critical accretion rate,
m ≤ mcrit, a stable, hot, optically thin, geometrically thick
solution can be found if radial energy transport (advection)
is included (see e.g. the review by Narayan 1997). The key
assumption of these models is that the protons gain the en-
ergy from viscous processes, while the electrons only acquire
energy through interactions (electron-ion coupling) with the
protons. This coupling timescale can be rather slow com-
pared to the accretion timescale, so protons can be accreted
into the black hole, taking some of the energy with them.
The energy that the electrons do manage to obtain is radi-
ated away via cyclotron/synchrotron emission (on an inter-
nally generated magnetic field), bremsstrahlung, or Comp-
ton scattering of the resultant spectra of these two processes.
In contrast to the SS accretion flow models, there is no cold
disk in the inner regions, so no strong source of soft seed
photons for Compton scattering, hence the resulting X -- ray
spectra are typically hard.
Such flows were proposed to explain the hard and very
faint X -- ray spectra seen from the transient X -- ray GBHC in
quiescence (e.g. Narayan, McClintock & Yi 1997), and then
extended by Esin, McClintock & Narayan (1997) to cover
the whole range of luminosity seen in the GBHC transients.
As the mass accretion rate increases from m << mcrit to
m ∼ mcrit the flow density increases, so the electron -- ion cou-
pling becomes more efficient and the fraction of energy the
electrons can drain from the protons increases. This increase
in radiative efficiency continues to m = mcrit, where only
∼ 35 % of the heat energy is advected. At higher m > mcrit
the advective flow collapses into an SS disk. Esin et al (1997)
identify this change from a hot advective flow to a cool SS
disk as the origin of the hard / soft spectral transition seen
in GBHC, and some observational support for this is indeed
seen ( Zycki et al., 1998a, Cui et al., 1998, Gierli´nski et al.,
1997b).
These models have been specifically applied to Cyg X -- 1
by Esin et al (1998). Their calculated spectra approximately
match the overall 2 -- 500 keV GINGA -- OSSE data from Cyg
X -- 1 for Rin ∼ 200Rg, but uncertainties in their model may
extend this to Rin ≥ 40Rg. This is still larger than most of
the constraints from relativistic smearing (table 2). Our data
indicate that the optically thick material extends much fur-
ther into the gravitational potential than anticipated by the
advective flows. Similar results are found for the comparison
of the models of Esin et al (1997) against the observed hard
state spectra from the GBHC transient source Nova Muscae
( Zycki et al., 1998a).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
6.3 Shapiro -- Lightman -- Eardley flows
An alternative X -- ray hot solution to the cool SS accre-
tion flow was discovered by Shapiro et al., (1976, hereafter
SLE). These solutions are actually very closely related to
the ADAFs discussed above in that they assume that the
protons gain most of the energy from viscous processes, and
then heat the electrons by electron -- ion coupling, resulting in
an optically thin(ish), geometrically thick(ish) flow. This so-
lution of the accretion flow equations merges into the ADAF
branch at a critical mass accretion rate (see Chen et al 1995;
Esin et al 1997; Zdziarski 1998). The key difference is that
on the SLE branch, Compton cooling dominates over advec-
tion, while the opposite is true on the ADAF branch. There
is no known possible hot solution for mass accretion rates
above the critical one at which the ADAF and SLE solutions
merge.
In its original form, the SLE solutions were shown to
be viscously stable but thermally unstable (Pringle 1976).
An increase in the heating rate is not balanced by an in-
crease in the cooling because the electron -- ion collisons be-
come less efficient as the temperature increases. The key
to the growth of interest in ADAFs is that they are sta-
bilsed against this due to the dominance of advective cool-
ing. However, the SLE is not necessarily thermally unstable.
The stability analyses which have been done generally as-
sumed that the viscous timescale is long compared to the
thermal timescale, so that the density of the region can-
not change during the perturbation. This is not true for a
geometrically thick(ish) flow. Many stability analyses also
neglected advective cooling, and/or external Compton cool-
ing on seed photons which vary. We urge further study of
the thermal stability of these flows.
Understanding the stability (and hence the existence) of
these flows is important since the SLE composite geometry
is rather similar to that inferred from the data, with a hot,
quasi -- spherical plasma cooled by radiation from an external
SS accretion disk. The advantage the SLE solution has over
the ADAF is that it radiates efficiently, which is generally re-
quired from the rapid observed variability of accreting black
holes. However, the disk transition again occurs where radi-
ation pressure dominates over gas pressure, so these models
again apparently predict that the inner disk radius is larger
at high mass accretion rates, opposite to the observed trend.
7 UNSTABLE ACCRETION IN AN SS DISK
The thermal instability above is produced wherever an in-
crease in the mass accretion rate is not balanced by a cor-
responding change in the cooling rate. Similarly, a disk is
viscously unstable if a small increase in the external mass ac-
cretion rate cannot be balanced by the viscosity to produce a
corresponding change in the mass accretion rate through the
disk. These two are related in that flows which are viscously
unstable are also necessarily thermally unstable.
The ionization of Hydrogen produces a disk transition
which is viscously (and thermally) unstable. Below temper-
atures of ∼ 5000 K the opacity is dominated by molecular
lines, while above this temperature there is a marked in-
crease in opacity due to bound -- free ionization of Hydrogen.
However, for temperatures beyond ∼ 104 K the bound -- free
12
C. Done & P.T. Zycki
opacity drops since hydrogen is then mostly ionized. This
gives a local maximum in opacity at ∼ 6000 − 104 K (e.g.
Hure et al 1994). In this range an increase in m gives rise to
a large increase in opacity, trapping the radiation and rais-
ing the central (equatorial plane) temperature of the disk.
This in turn raises the sound speed, cs, and the kinematic
viscosity, ν = αcsH, leading to an increase in the rate at
which the disk material is accreted onto the compact object
(see e.g. the review by Osaki 1996).
If this instability can propagate throughout the disk
then this leads to two distinct states. In quiescence the disk
is cool and has low kinematic viscosity -- generally so low
that the rate at which material is added to the disk from
the external accretion source is faster than the rate at which
mass can be viscously transported inwards from the outer
disk. Mass builds up at the outer edge of the disk, and the
structure is not that of a cool steady SS disk. Eventually, the
outer disk mass is so large that hydrogen can be ionized, and
then the disk switches to a hot state where the kinematic
viscosity is high and steady state can be achieved.
This instability is thought to be the trigger for the out-
burst in the transient black hole systems (see e.g. Lasota,
Narayan & Yi 1996), though it was first discovered in the
context of the quiescent/outburst behaviour of disk accret-
ing white dwarfs (see e.g. the review by Osaki 1996). How-
ever, one major problem with these models is that the ac-
cretion rate inferred in quiescence (as evinced by the X --
ray emission) is rather larger than the maximum accretion
rate for which the cool quiescent disk can be sustained if
the disk extends all the way down to the compact object
(Lasota et al., 1996). This led to the idea of a 'hole' in
the inner disk in quiescence, perhaps formed by evaporation
(Meyer & Meyer -- Hofmeister 1994). However, another possi-
bility which has yet to be explored in this context is that the
inner disk in quiescence is simply optically thin. The surface
density profile for the non -- stationary cool quiescent disk is
peaked to the outside of the disk, so it is very easy for the
inner regions to be optically thin (e.g. Cannizzo 1998, figure
7), especially as the absorption opacity of the cool disk is low
so even if the photon is scattered many times, there is still
a high probability that it can ultimately escape (effectively
optically thin). This means that the inner disk is unlikely
to be able to thermalise the energy released locally, and so
may not radiate (i.e. cool) efficiently. The material would
then heat up, and turn into something like an advective or
externally cooled SLE flow which would appear to emanate
from a (non stationary) optically thick disk.
However, the main problem with such models is that
while this may describe the situation in quiescence, it seems
highly unlikely that it has any relevance to the high mass
accretion rates required for Cyg X -- 1 and the other X -- ray
bright transient GBHC. Another instability that can give
the same effect but at high mass accretion rates is required.
One obvious possibility is the instability which occurs when
radiation pressure starts to dominate in the disk. The in-
ternal pressure (and hence the sound speed and kinematic
viscosity) then rises very rapidly (∝ T 4), and the mass ac-
cretion rate through the radiation pressure dominated part
of the disk can then increase faster than the increase in
external mass accretion which triggered the rise in temper-
ature, so leading to depletion of the inner disk. At higher
mass accretion rates the flow is again stabilised though the
excess radiation being carried along with the flow and ad-
vected into the black hole (Abramowicz et al 1988). Models
of such flows are beginning to be developed (Szuszkiewicz &
Miller 1998) in response to observations of rapid inner disk
instabilities in the superluminal accreting black hole candi-
date GRS1915+105 (Belloni et al 1997), but are currently
in their infancy. It remains to be seen whether models of a
non -- stationary flow inside the radiation pressure dominated
regime could reproduce what is seen in Cyg X -- 1 and other
low state GBHC spectra.
8 INHOMOGENEOUS ACCRETION
It seems highly unlikely that accretion flow can change
smoothly between an optically thick cool solution and an
optically thin hot solution. Some transition region where
cool clumps are embedded in hot plasma seems physically
more probable, and turbulence is indeed seen in the nu-
merical calculations of Abramowicz, Igumenshchev & La-
sota (1998). Recent work on an inhomogeneous solution to
the accretion flow equations has shown that there is a sta-
ble solution to the radiation pressure dominated SS disk
if the majority of mass is in dense clumps which are opti-
cally thick, while the remainder forms an optically thin(ish),
X -- ray hot plasma (Krolik 1998), with a physical mecha-
nism for clumping provided by the photon bubble instability
(Gammie 1998). However, one problem with these models is
that they seem to predict that the covering fraction of cool
clumps is close to unity for stellar mass black holes, in con-
flict with the observations.
9 CONCLUSIONS
We detect significant relativistic smearing of the reflected
spectral features in Cyg X -- 1, giving the first observational
confirmation of an inner accretion disk in Cyg X -- 1. Our
model calculates the total (line plus continuum) reflected
spectrum for a given ionization state and iron abundance
and then applies the relativistic smearing at a given incli-
nation. These four parameters can be uniquely determined
from X -- ray spectra (assuming a given illuminating flux)
with moderate energy resolution, and excellent signal -- to --
noise data extending beyond 9 keV. These conditions are
satisfied in Cyg X -- 1 for the EXOSAT GSPC and ASCA GIS
spectra. The GINGA data are taken in a reduced spectral
resolution mode, while the ASCA SIS data have insufficient
signal -- to -- noise above 9 keV.
For the ASCA GIS and EXOSAT GSPC data the re-
flected spectrum is constrained to be from material inclined
at ≥ 45◦ which is not highly ionized. The low ionization state
rules out Auger ionization as a mechanism for suppressing
the line with respect to the reflected continuum as proposed
by Ross et al (1996). The iron line is not weak with respect
to the reflected continuum -- we infer twice solar iron abun-
dance -- but the visibility of the line is dramatically reduced
by relativistic smearing.
The observed amount of relativistic smearing is not con-
sistent with a disk extending all the way down to the last
stable orbit around the black hole, but instead indicates that
the disk truncates at some typical (but probably variable)
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Relativistic Distortions in the X -- ray Spectrum of Cyg X -- 1
13
radius of ∼ 10 − 30Rg, with a trend that larger transi-
tion radii are associated with harder X -- ray spectra. This
mirrors the results from the hard spectra of the transient
GBHC such as V404 Cyg and Nova Muscae ( Zycki et al
1997; 1998ab), suggesting that low state can be associated
with a lack of optically thick disk at the innermost radii.
The amplitude of the reflected component is signifi-
cantly less than that expected from an isotropic source above
a semi -- infinite disk. Again this result can be understood in
the context of an inner 'hole' in the optically thick material
which is filled by the X -- ray emitting plasma. Similar con-
clusions about the geometry are also found from looking at
the detailed continuum spectral shape of Cyg X -- 1, where
the hardness of the X -- ray spectrum implies that the hot X --
ray plasma is spatially separate from most of the accretion
disk, otherwise it intercepts too many soft photons from the
disk and produces an X -- ray spectrum which is significantly
steeper than observed (Gierli´nski et al 1997a; Dove et al
1997; Poutanen et al 1997).
The data then strongly indicate that low state GBHC
accretion geometry is that of an inner X -- ray hot, optically
thin(ish), geometrically thick plasma, which cools by Comp-
ton upscattering of soft photons from an external cool, op-
tically thick, geometrically thin disk. Such a composite flow
seems most likely to be a standard SS cool outer disk, with
an inner ADAF or SLE flow. The SLE flow is probably
thermally unstable, so for steady state solutions the ADAF
seems to be preferred. However, we stress that the radiation
pressure instability may put the disk in a regime where the
flow is not in steady state, or is inhomogeneous.
10 ACKNOWLEDGEMENTS
We thank Ken Ebisawa and Yoshihiro Ueda for kindly al-
lowing us to use their Cyg X -- 1 ASCA data, Andrew King,
Uli Kolb and Juri Poutanen for stimulating discussions, and
Andrzej Zdziarski the referee for useful comments. CD ac-
knowledges support from a PPARC Advanced Fellowship.
REFERENCES
Abramowicz M.A., Igumenshchev I.V., Lasota J. -- P., 1998, MN-
RAS, 293, 443
Abramowicz M.A., Czerny B., Lasota J. -- P., Szuszkiewicz E.,
1988, ApJ., 332, 646
Ebisawa K., Ueda Y., Inoue H., Tanaka Y., White, N.E. 1996,
ApJ., 467, 419
Esin A. A., McClintock J.E., Narayan R., 1997, ApJ., 489, 865
Esin A.A., Narayan R., Cui W., Grove J.E., Zhang S -- N., 1998,
ApJ., 505, 854
Fabian A.C., Rees M.J., Stella L., White, N.E. 1989, MNRAS,
238, 729
Frank J., King A.R. & Raine D., 1992, In "Accretion power in
Astrophysics ", 2d edition, Cambridge, UK: Cambridge Uni-
versity Press
Gammie C., 1998, MNRAS, 297, 929.
George I.M., Fabian A.C., 1991, MNRAS, 249, 352
Gierli´nski M., Zdziarski A. A., Done C., Johnson W. N., Ebisawa
K., Ueda Y., Phlips F. 1997a, MNRAS, 288, 958
Gierli´nski M., Zdziarski A.A., Dotani T., Ebisawa K., Jahoda K.,
Johnson W.N., 1997b, in Proc. 4th Compton Symposium, eds
Dermer C.D., Strickman M.S., Kurfess J.D., AIP 410, p844
Gies D.R., Bolton C.T., 1986, ApJ., 304, 371
Haardt F., Maraschi L., 1993, ApJ., 413, 507
Haardt F., Done C., Matt G., Fabian A.C., 1993, ApJL., 411, 95
Haardt F., Maraschi L., Ghisellini G., 1994, ApJ., 432, 95
Hure, J. -- M., Collin -- Souffrin S., Le Bourlot J., Pineau des Forets
G., 1994, A& A., 290, 34
Iwasawa K., Fabian A.C., Mushotzky R.F., Brandt W.N., Awaki
H., Kunieda H., 1996, MNRAS, 279, 837
Kaastra J.S., Mewe R., 1993, A&AS, 97, 443.
Kitamoto S., Miyamoto S., Tanaka Y., Ohashi T., Kondo Y.,
Tawara Y., Nakagawa M., 1984, PASJ, 36, 799
Krolik J.H., 1998, ApJL., 498, 13
Lasota J. -- P., Narayan R., Yi I., 1996, A&A, 314, 813
Lightman A.P., White T.R., 1988, ApJ, 335, 57
Magdziarz P. Zdziarski A.A., 1995, MNRAS, 273, 837
Marshall F.E. Mushotzky R.F., Petre R., Serlemitsos P.J., 1993,
ApJ., 419, 301
Matt G., Perola G.C., Piro L. 1991, A&A, 247, 25
Matt G., Fabian A.C., Ross R., 1993a, MNRAS, 262, 179.
Matt G., Fabian A.C., Ross R., 1993b, MNRAS, 264, 839
Matt G., Fabian A.C., Ross R., 1996, MNRAS, 278, 1111
Meyer F., Meyer -- Hofmeister E., 1994, A&A, 288, 175
Morrison R., McCammon D., ApJ., 270, 119
Nandra K., Pounds K.A., 1994, MNRAS, 268, 405
Narayan R. 1997, in Proc. IAU Colloq. 163 on Accretion Phe-
nomena and Related Outflows, eds. D. T. Wickramasinghe,
L. Ferrario, G. V. Bicknell, p. 75
Narayan R., McClintock J. E., Yi I., 1996, ApJ, 457, 821
Narayan R., Yi I., 1995, ApJ., 444, 231
Osaki Y., 1996, PASP, 108, 39
Paciesas W.S., Robinson C.R., McCollough M.L., Zhang S.N.,
Harmon B.A., Wilson C.A., 1997, in Proc. 4th Compton Sym-
posium, eds Dermer C.D., Strickman M.S., Kurfess J.D., AIP
410, p834
Abramowicz M.A., Chen X., Kato S., Lasota J. -- P., Regev O.,
Pounds K.A., Nandra K., Stewart G.C., George I.M., Fabian
1995, ApJL., 438, 37
Balucinska M., Hasinger G., 1991, A&A, 241, 439
Basko M.M., 1978, ApJ., 223, 268
Belloni T., Mendez M., King A.R., Van Der Klis M., Van Paradijs
J., ApJL., 1997, 488, 109
Blondin J., Stevens I., Kallman T.R., 1991, ApJ., 371, 684
Cannizzo J., 1998, ApJ., 494, 366
Chen X., Abramowicz M.A., Lasota J. -- P., Narayan R., Yi, I.,
ApJL, 1995 443, 61
Cui W., Ebisawa K., Dotani T., Kubota, A., 1998, ApJ., 493, 75
Di Salvo T., Done C., Zycki P., Burderi L., Robba N., In 'The Ex-
treme Universe', Ed Winkler C., Gordon & Breach, in press.
Done C., Mulchaey J.S., Mushotzky R.F., Arnaud K.A., 1992,
ApJ., 395, 275
Dove J., Wilms J., Maisack M., Begelman M.C., 1997, ApJ., 487,
759
A.C., 1990, Nature, 344, 132
Poutanen J., Krolik J.H., Ryde F., 1997, MNRAS, 292, L21
Poutanen J., 1998, In 'Theory of Black Hole Accretion', CUP, in
press.
Pringle J., 1976, MNRAS, 177, 65
Reilman R.F., Manson S.T., 1979, ApJS., 40, 815
Ross R.R., Fabian A.C., 1993, MNRAS, 261, 74.
Ross R.R., Fabian A.C., Brandt W.N., 1996, MNRAS, 278, 1082
Shakura, N.I., Sunyaev, R.A. 1973, A& A, 24, 337
Shakura, N.I., Sunyaev, R.A. 1976
Shapiro S.L., Lightman A.P., Eardley D.M., 1976, ApJ, 204,187
Sowers J.W., Gies D.R., Bagnuolo W.G., Shafter A.W., Wiemker
R., Wiggs M.S., 1998, ApJ., 505, 424
Szuskiewicz E., Miller J.C., 1998, MNRAS, 298, 888
Tanaka Y. et al. 1995, Nature, 375, 659
Zdziarski A.A., 1998, MNRAS, 296, L51
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
14
C. Done & P.T. Zycki
Zdziarski A.A., Lubi´nski P., Smith D.A., 1999, MNRAS, in press
Zycki P.T., Done C., Smith D.A., 1997, ApJL, 488, 113
Zycki P.T., Done C., Smith D.A., 1998a, ApJL, 496,25
Zycki P.T., Done C., Smith D.A., 1998b, MNRAS, submitted
Zycki P.T., Czerny, B. 1994, MNRAS, 266, 653
Zycki P.T., Krolik J.H., Zdziarski A.A., Kallman T.R., 1994,
ApJ., 437, 597.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/9411019 | 1 | 9411 | 1994-11-04T22:04:23 | Theory of Exploring the Dark Halo with Microlensing 1: Power--Law Models | [
"astro-ph",
"hep-ph"
] | The detection of microlensing has opened the way for the development of new methods in galactic astronomy. This series of papers investigates what microlensing can teach us about the structure and shape of the dark halo. In this paper we present formulas for the microlensing rate, optical depth and event duration distributions for a simple set of axisymmetric disk-halo models. The halos are based on the "power--law models" which have simple velocity distributions. Using these models, we show that there is a large uncertainty in the predicted microlensing rate because of uncertainty in the halo parameters. For example, models which reproduce the measured galactic observables to within their errors still differ in microlensing rate towards the Magellanic Clouds by more than a factor of ten. We find that while the more easily computed optical depth correlates well with microlensing rate, the ratio of optical depth to rate can vary by a factor of two (or greater if the disk is maximal). Comparison of microlensing rates towards the Large and Small Magellanic Clouds (LMC and SMC) and M31 can be used to aid determinations of the halo flattening and rotation curve slope. For example, the ratio of microlensing rates towards the LMC and SMC is $\sim 0.7-0.8$ for E0 halos and $\sim 1.0 - 1.2$ for E7 halos (c.f. Sackett \& Gould 1993). Once the flattening has been established, the ratio of microlensing rates towards M31 and the LMC may help to distinguish between models with rising, flat or falling rotation curves. Comparison of rates along LMC and galactic bulge lines-of-sight gives useful information on the halo core radius, although this may not be so easy to extract in practice. Maximal disk models provide substantially smaller halo optical depths, shorter event durations and even larger model uncertainties. | astro-ph | astro-ph |
THEORY OF EXPLORING THE DARK HALO
WITH MICROLENSING 1: POWER–LAW MODELS
C. Alcock∗,† , R.A. Allsman‡ , T.S. Axelrod‡ , D.P. Bennett∗,†,
K.H. Cook∗,† , N.W. Evans♣ , K.C. Freeman‡ , K. Griest†,k ,
J. Jijinak , M. Lehnerk, S.L. Marshall†,♭ , S. Perlmutter†,
B.A. Peterson‡, M.R. Pratt†,♭ , P.J. Quinn‡, A.W. Rodgers‡ ,
C.W. Stubbs†,♭, W. Sutherland♠
(The MACHO Collaboration)
∗ Lawrence Livermore National Laboratory, Livermore, CA 94550
† Center for Particle Astrophysics, University of California, Berkeley, CA 94720
‡ Mt. Stromlo and Siding Spring Observatories,
Australian National University, Weston, ACT 2611, Australia
♣ Theoretical Physics, Department of Physics, University of Oxford, OX1 3NP, UK
k Department of Physics, University of California, San Diego, CA 92039
♭ Department of Physics, University of California, Santa Barbara, CA 93106
♠ Astrophysics, Department of Physics, University of Oxford, OX1 3RH, UK
Submitted to the Astrophysical Journal, 4 November, 1994
Abstract
If microlensing of stars by dark matter has been detected (Alcock et al.
1993; Aubourg et al. 1993; Udalski et al. 1993; Alcock et al. 1994; Udalski et al.
1994a,b), then the way is open for the development of new methods in galactic
astronomy. This series of papers investigates what microlensing can teach us
about the structure and shape of the dark halo. In this paper we present formulas
for the microlensing rate, optical depth and event duration distributions for a
simple set of axisymmetric disk–halo models. The halos are based on the “power–
law models” (Evans 1993, 1994) which have simple velocity distributions.
Using these models, we show that there is a large uncertainty in the predicted
microlensing rate because of uncertainty in the halo parameters. For example,
models which reproduce the measured galactic observables to within their errors
still differ in microlensing rate towards the Magellanic Clouds by more than
a factor of ten. We find that while the more easily computed optical depth
correlates well with microlensing rate, the ratio of optical depth to rate can vary
by a factor of two (or greater if the disk is maximal). Comparison of microlensing
rates towards the Large and Small Magellanic Clouds (LMC and SMC) and
M31 can be used to aid determinations of the halo flattening and rotation curve
1
slope. For example, the ratio of microlensing rates towards the LMC and SMC is
∼ 0.7 − 0.8 for E0 halos and ∼ 1.0 − 1.2 for E7 halos (c.f. Sackett & Gould 1993).
Once the flattening has been established, the ratio of microlensing rates towards
M31 and the LMC may help to distinguish between models with rising, flat or
falling rotation curves. Comparison of rates along LMC and galactic bulge lines-
of-sight gives useful information on the halo core radius, although this may not be
so easy to extract in practice. Maximal disk models provide substantially smaller
halo optical depths, shorter event durations and even larger model uncertainties.
Sub ject headings: dark matter - Galaxy: structure - gravitational lensing
1. Introduction
The recent detection of possible gravitational microlensing events (Alcock et
al. 1993; Aubourg et al. 1993; Udalski et al. 1993, Alcock et al. 1994; Udalski et al.
1994a,b) gives hope that at least part of the dark matter content of our galaxy is
directly accessible to observation. The dark halo, whose extent has been studied
gravitationally for many years via velocities of stars, gas, and satellites (e.g. Fich
& Tremaine 1991), contains at least three times (and perhaps more than ten
times) the mass of the luminous galaxy. Its identity is one of the ma jor unsolved
problems in astronomy (e.g. Ashman 1992; Primack, Seckel, & Sadoulet 1988).
While it is possible that the halo consists mostly of exotic non-baryonic elemen-
tary particles, the idea of Paczy´nski (1986) of searching for Massive Compact
Halo Ob jects (Machos) in the range 10−8M⊙ to 103M⊙ by monitoring millions
of stars in the LMC may have borne fruit in the experimental programs.
If the Milky Way halo contains large numbers of Machos, then the gravi-
tational microlensing experiments now under way should have the potential to
determine the number and distribution of Machos in the halo, as well as their
mass distribution. The next step for the microlensing experiments will be to
gather more events and then translate the number and duration of those events
into an estimate of the mass fraction f of the dark halo which consists of Machos
in the relevant mass range. To accomplish this goal a model of the dark halo
is necessary. In the past simple spherical models with flat rotation curves have
been considered (Paczy´nski 1986; Griest 1991; DeRujula, Jetzer & Masso 1991;
Nemiroff 1991). These have been valuable in estimating the order–of–magnitude
effects but suffer from at least three important deficiencies:
(1) The halo may not be spherical. N-body simulations of gravitational collapse
of collisionless dark matter generically produce axisymmetric or triaxial halos
(Quinn, et al. 1992; Dubinski & Carlberg 1991; Katz 1991). The recent papers
of Sackett & Gould (1993) and Frieman & Scoccimarro (1994) have made an
important start on the study of microlensing effects in flattened halos (see also
the early work of Jetzer 1991).
(2) The effect of the galactic disk is ignored. The disk makes a significant con-
tribution to the local circular speed. Modeling without proper allowance for this
effect leads us to over–estimate the gravity field and hence the mass of the dark
halo. That is, since the amount of material in the galactic halo is set by the local
circular speed, a larger contribution to this speed by the disk, means a smaller
halo is needed to explain the total circular velocity.
(3) It is only a simplified view of the data that permits one to regard the rotation
curve of the Milky Way as flat. In fact, even the sign of the local gradient of the
rotation law at the sun is not known – Fich, Blitz & Stark (1989) estimate that
2
it may be rising or falling by about 30 km/sec outwards from the solar circle to
a galactocentric radius of 17 kpc.
In this paper, we make a start towards quantifying and remedying these defects
by calculating the microlensing rate and optical depth in a set of simple, flexible
and realistic halo models. We take the power–law models (Evans 1993, 1994,
hereafter E93, E94) – for which simple and self-consistent distribution functions
are known – and provide an approximate method to allow for the influence of the
galactic disk. This enables calculation not just of the optical depth, but also the
event rate and the distribution of event durations. In this paper, the emphasis is
on the uncertainties in microlensing predictions and what can be done to reduce
them. Since there are many possible models of the dark halo consistent with
current observations, there is a substantial scatter in the predicted microlensing
rate – and this in turn contributes to an uncertainty in the measurement of the
fraction of the halo consisting of Machos. Of course, this is directly relevant to
the difficult but important question of whether the observed microlensing rates
can rule out or support the existence of non–baryonic matter in the halo.
We show that the often calculated optical depth is a useful predictor of the
microlensing rate to a factor of ∼ 2. For more accurate work and to predict event
durations, a distribution of Macho velocities is needed. By considering a large
range of model parameters consistent with observations, we find the microlensing
optical depth can vary by a factor of six or more.
If the disk of the Milky
Way is “maximal” – in the sense that it provides almost all of the Galacto-
centric acceleration – then a much smaller halo is required. This gives an even
larger spread in predicted optical depth, rate, and event durations. An attractive
possibility is to use the measured microlensing rates toward different sources (e.g.
LMC, SMC, M31 and the galactic bulge) to determine some of the halo and disk
parameters, thereby providing a new tool for the study of galactic structure and
at the same time reducing the halo uncertainty in the measurement of f . We
corroborate the Sackett & Gould (1993) prediction that the ratio of LMC to
SMC optical depth is a robust indicator of the flattening of the dark halo – and
extend it by showing that the ratio of microlensing rates distinguishes flatness as
well. We predict that the ratio of rates toward M31 and the LMC may be enable
us to discover whether the rotation curve is rising or falling (or equivalently the
extent of the dark halo). A comparison of the bulge and LMC rates can provide
information on the halo core radius.
The plan of the paper is as follows: In §2 we review the axisymmetric models
and present formulas for the optical depth, microlensing rate, and event dura-
tions. In §3 we discuss the halo parameters and their allowed range and explain
how to take into account the effect of the galactic disk. In §4 we compare op-
tical depth and rate, and discuss the uncertainties in microlensing rates due to
uncertainties in the parameters. In §5 we discuss reducing those uncertainties by
comparing results along different lines-of-sight, in §6 we discuss distributions of
event durations, and in §7 we summarize our conclusions.
2. Axisymmetric models
The primary goal of the galactic gravitational microlensing experiments is
to determine the mass of the dark halo in Machos. The experiments search for
Machos by monitoring millions of stars nightly in the Large Magellanic Cloud
(LMC), the Small Magellanic Cloud (SMC), the galactic bulge, and perhaps in the
future in the M31 galaxy (Crotts 1992; Baillon et al. 1993). If the dark halo con-
tains large numbers of Machos, occasionally one passes close to the observer–star
line–of–sight and acts as a gravitational lens, causing a time-dependent magnifi-
cation of the stellar image. The resulting lightcurve is determined by only a few
3
quantities such as the distance s from us to the Macho, Macho mass m, Macho
transverse velocity v⊥ , and impact parameter b. The magnification A(t) as a
function of time t is given by
A(t) = (u2 + 2)/[u(u2 + 4)1/2 ],
min + ω 2(t − t0 )2)1/2 ,
u(t) = b/Re = (u2
Re = (2/c)[Gms(1 − s/L)]1/2 ,
(1)
where the peak magnification Amax is given by inverting umin = u(Amax), ω =
v⊥/Re , and L is the distance to the star. Experimentally, microlensing events
are characterized by the maximum magnification Amax , the time of the peak t0 ,
and the duration of the event bt, where bt = 2/ω . Also used in the literature as an
“event duration” is te = bt(u2
T − u2
min )1/2 . This is time for which A ≥ AT , with
AT = A(uT ). Using AT = 1.34 corresponds to uT = 1 which is the time inside
the Einstein radius Re . A more thorough discussion can be found in in Paczy´nski
(1986) and Griest (1991).
An observing team measures the number and duration of microlensing events.
The number of observed events is proportional to the number of stars monitored,
the duration of the experiment, the experimental efficiency, and the rate at which
microlensing occurs. The primary observables are the optical depth τ , the rate
Γ and the average duration of the events (cid:10)bt(cid:11). They are related by
Γ (cid:10)bt(cid:11) .
π
τ = Γ hte i =
4
If the distribution of Machos masses n(m) were a delta function, then Γ would
be Γ = Γ1 (m/M⊙ )−1/2 , where Γ1 is the rate with m = M⊙ and is independent
of mass. For a general normalized mass distribution Γ = ηmΓ1 , where the mass
integral is
ηm = Z dmn(m)(m/M⊙ )−1/2 .
This also implies that (cid:10)bt(cid:11) = (cid:10)bt(cid:11)1 η−1
m .
The optical depth is the number of Machos inside the microlensing tube with
radius uT Re(s) and length L. It depends only on the density of Machos ρ
τ = Z ρ(x)
m
Unlike the rate or average duration, it is independent of the Macho mass distri-
bution. For this reason it also contains no direct information about the mass of
the lensed ob jects. Note also that since detection efficiencies depend upon the
duration of events, it is important to have models which predict durations. Now
to find the rate at which Machos enter the microlensing tube requires knowledge
πu2
T R2
e (s)ds.
(4)
(2)
(3)
4
of the distribution of velocities all along the tube. So the rate and the distri-
bution of event durations are hard to calculate because they require the entire
phase space distribution function (DF) F (v, x). The differential rate is given by:
dΓ =
1
m
F (v, x) cos θuT Rev⊥d3vdxdα,
(5)
where the angles and notation are defined in Griest (1991).
What makes the calculation particularly difficult is that the DF cannot be
prescribed arbitrarily as a Maxwellian, for instance. This is because the Machos
are collisionless, so the DF is constrained to obey the collisionless Boltzmann
equation. By Jeans’ theorem, this implies that the DF depends only on the iso-
lating integrals of motion (see Binney & Tremaine 1987, p. 220). Self–consistent
solutions for distributions of velocities that build flattened halo models are scarce.
The largest known set of axisymmetric models with simple DFs are the “power–
law galaxies” (E93, E94). These form the basis for the exploration of microlensing
in this paper, allowing us to go beyond simple spherical models. Note, however,
that all these models are axisymmetric and oblate, while N-body simulations
suggest that halos may well be triaxial. Exploration of triaxial models will be
done in a future paper.
The parameters of the power–law models are:
(1) The core radius Rc , which measures the scale at which the density law begins
to soften.
(2) The flattening parameter q , which is the axis ratio of the concentric equipo-
tential spheroids, with q = 1 representing a spherical (E0) halo and q ∼ 0.7
representing an ellipticity of about E6. The “isophotal” ellipticity of the dark
halo is a function of q , as well as other parameters of the model [see E94, eq.
(2.9)].
(3) The parameter β , which determines whether the rotation curve asymptotically
rises, falls or is flat. At large distances R in the equatorial plane, the rotation
velocity vcirc ∼ R−β . So β = 0 corresponds to a flat rotation curve, while β < 0
is a rising rotation curve and β > 0 is falling.
(4)The solar radius R0 , which is the distance of the Sun from the galactic center.
(5) Finally, the normalization velocity v0 , which determines the overall depth
of the potential well and hence the typical velocities of Machos in the halo. In
the limit β = 0, q = 1 and large R (spherical halo with a flat rotation curve)
v0 = vcirc .
Using z as the height above the equatorial plane, the potential of the power–
law models is
0 Rβ
v2
c /β
c + R2 + z2 q−2)β/2
(R2
log(R2
c + R2 + z2 q−2 ),
v2
0
2
−
,
if β 6= 0,
Ψ =
(6)
if β = 0,
5
and the mass density is
F (E , Lz ) =
if β 6= 0,
ρ =
.
(7)
z E 4/β−3/2 + B E 4/β−1/2 + C E 2/β−1/2 ,
AL2
c (1 + 2q2 ) + R2(1 − β q2 ) + z2 (2 − (1 + β )q−2 )
R2
c + R2 + z2 q−2)(β+4)/2
(R2
0 Rβ
v2
c
4πGq2
The DF corresponding to this potential–density pair is
if β = 0.
(8)
where the constants A, B , and C are given in E93 and E94. As required by Jeans’
theorem, the DFs depend only on the isolating integrals of motion, namely the
relative energy per unit mass E = Ψ − 1
2 v2 , and the angular momentum per unit
mass about the symmetry axis Lz . The circular velocity in the equatorial plane
is
0 ) + C exp(2E /v2
0 ) + B exp(4E /v2
z exp(4E /v2
AL2
0 ),
v2
circ =
0 Rβ
c R2
v2
c + R2)(β+2)/2
(R2
.
(9)
Note that the limit q = 1, β = 0 and Rc = 0 recovers the standard singular
isothermal sphere used by Paczy´nski. Allowing a core radius gives
ρ =
v2
0
4πG
R2 + 3R2
c
c )2 .
(R2 + R2
(10)
This differs from the cored isothermal sphere considered by Griest (1991) in sev-
eral ways. First, the rotation curve approaches its asymptotic value more quickly.
Second, the DF given by the q = 1, β = 0 limit of equation (8) is self–consistent,
whereas Griest (1991) assumed an approximate Maxwellian distribution of ve-
locities.
So far we have only modeled the dark halo. However, in the standard model,
a substantial fraction (∼ 40%) of the centripetal force at the solar radius derives
from the disk stars. This is represented by a thin exponential disk with a scale
length of Rd = 3.5 kpc, normalized to a surface density of Σ0 = 50M⊙pc−2 at the
solar radius (Gilmore et al. 1989; Gould 1990). It is possible that the disk of our
galaxy is substantially larger than the canonical value. (Oort 1960; Bahcall 1984;
Kuijken and Gilmore 1989; Gould 1990). Recent microlensing results (Alcock,
et al. 1994; Udalski, et al. 1994) as well as studies of the optical rotation curves
of external galaxies (Buchhorn 1992; Kent 1992) may suggest this. We consider
such a “maximal disk” by taking Σ0 = 100M⊙pc−2 . The rotation velocity added
in quadrature is thus (Freeman 1970; Binney & Tremaine 1987, p. 77)
v2
disk = 4πGΣ0hy 2 [I0 (y )K0(y ) − I1 (y )K1(y )] ,
where y = R/(2Rd), and the In and Kn are modified Bessel functions. Note
that in adding a contribution from the disk to the local circular velocity, we have
(11)
6
sacrificed self–consistency. Really, we should find the DF of the power–law halo
in the combined potential field of both disk and halo – instead, we use the DF
(8). As has been argued elsewhere (Evans & Jijina, 1994), this is an reasonable
approximation for the LMC, SMC and M31, where microlensing typically occurs
at heights above the equatorial plane of many kpc.
In this paper, our aim is to estimate the contribution of the Galactic halo to
microlensing. Of course, this is not the only possible source of deflectors. Towards
the LMC, there is the possibility of microlensing by the LMC dark halo or disk
(Gould 1993b, Sahu 1994). The optical depth is ∼ 2.5 × 10−7 for microlensing by
LMC halo lenses and ∼ 0.09× 10−7 for LMC disk lenses. The Galactic halo makes
a contribution that is roughly three times greater and so is the dominant source
of lenses. However, this is not the case for lines of sight towards M31. The optical
depth is dominated by Machos in the halo and disk of M31 (Crotts 1992, Gould
1993a). Crotts (1992) estimates that the halo of our own Galaxy contributes just
20% to the total optical depth. Microlensing towards the Galactic bulge poses
perhaps the hardest problems of separating the contributions of different deflector
populations. Bulge stars can undergo microlensing not only by halo Machos, but
also by other bulge and disk stars (Griest et al. 1991, Paczy´nski 1991, Kiraga
& Paczy´nski 1994). At Baade’s Window, the optical depth is ∼ 6.3 × 10−7 for
microlensing by bulge lenses, ∼ 5.0 × 10−7 for disk lenses. The dark halo only
makes an important contribution if the core radius is small.
We are now in a position to calculate the microlensing observables – the
optical depth, rate and average duration of events. They can be found using
equations 2, 3, 4 and 7. The results are single quadratures and readily evaluated
on the computer. They are displayed in Appendix A. In Appendix B, we give
the differential microlensing rate dΓ/dbt, where bt is defined just after equation (1).
The probability of obtaining an event of duration bt is just (dΓ/dbt)/Γ.
3. Range of models
In order to explore the scatter in microlensing observables, we build a set
of halo models which span the observationally allowed range. The power–law
galaxy models allow us to vary the flattening, core radius and rotation law, and
we consider both canonical and maximal disks. For each parameter in the model,
we therefore find the range permitted by the observations. Then, several values
of each parameter are chosen to represent the range. We also ensure that each set
of parameters gives a model consistent with the measured Milky Way rotation
curve. So, we study the statistical properties of an ensemble of models, each one
of which is a plausible representation of the dark halo of the Milky Way.
For the dark halo flattening, little is known. So the entire range of flatten-
ing allowed by the power–law models is examined. This varies between E0 or
spherical (q = 1) and roughly E6 or E7 (depending on β ). The core radius of the
dark halo is also uncertain – Bahcall, Schmidt & Soneria (1983) estimate Rc as
2 kpc from star count data, while Caldwell & Ostriker (1981) suggest 10 kpc. If
the disk is maximal, values as large as 20 kpc are possible. We consider values
of 2 kpc, 5 kpc, 10 kpc, and 20 kpc. The parameter β determines the slope of
asymptotic circular velocity. Between R0 and 2R0 , the circular velocity is prob-
ably within 10 − 15% of the I.A.U value of 220 km/s, but whether the measured
HI rotational velocities rise or fall with R depends upon estimates of the solar
position R0 and the local circular speed vcirc (R0) (see Fich et al. 1989; Jones et
al. 1993). Beyond 20 kpc, little is known directly, though arguments based on
the kinematics of distant satellite galaxies support the idea of a relatively flat
7
rotation curve out to an unknown cut–off (Fich & Tremaine 1991). However,
current theories of galaxy formation tend to favor the alternative view that dark
halos extend indefinitely, fading into structure on larger scales. So, we do not
consider a halo cut–off in this paper – it would add yet another poorly known
parameter to our model. We investigate power–law halos with β = -0.2, 0, and
0.2. These correspond to rotation curves which rise by ∼15%, are flat, or fall
by ∼15% between the solar radius and twice the solar radius, depending a little
upon Rc .
The value of the solar radius R0 has been reviewed by Reid (1989). He
shows that most recent determinations lie between 7 kpc and 9 kpc, with 7.7
kpc being his preferred value. This differs considerably from the IAU value of
8.5 kpc (Kerr & Lynden–Bell 1986). We examine the values R0 = 7, 8, and 9
kpc. Finally, perhaps the single most important parameter is the normalization
velocity v0 . Given our fixed disk contribution to the total rotation law, the
parameter v0 is now specified once we settle upon a choice for vcirc (R0). Merrifield
(1992) estimates vcirc (R0) = 200 ± 10 km/s, Fich, Blitz, & Stark (1989) give
vcirc (R0) = 220 ± 30 km/s, while Rohlfs et al. (1986) give values between 170
km/s and 200 km/s between R0 = 6 kpc and R0 = 16 kpc. For our ensemble of
models, we impose the constraint that the total circular velocity lies between 180
km/s and 250 km/s at R0 and 2R0 . Note that the IAU value is 220 km/s (Kerr
& Lynden-Bell 1986). We also investigated a more restricted ensemble of models
with 190 ≤ vcirc (R0) ≤ 230 km/s. We find all our results also hold for this more
restricted ensemble.
4. Uncertainties in the Rates
First, let us consider the difference caused by using the optical depth instead
of the microlensing event rate. The optical depth to microlensing is the mean
number of Machos in the microlensing tube; that is the number of microlensing
events taking place at a given moment. It is easy to calculate since it is inde-
pendent of lens mass and velocity, and only requires knowledge of the density
distribution ρ(x). For this reason, it is the most widely estimated quantity. But
how well does it trace the microlensing rate?
We are able to answer this question since both the rate Γ (equation A1) and
the optical depth τ (equation A6) are known for the power–law models. One way
to test this is to plot (cid:10)bt(cid:11), which is the ratio of optical depth τ and Γ, (cid:10)bt(cid:11) = 4
π τ /Γ,
for many different models. The average duration (cid:10)bt(cid:11) is a constant if τ and Γ are
well–correlated. In Fig. 1, we show histograms of (cid:10)bt(cid:11) for microlensing towards
the LMC, SMC and M31 for our ensemble of models. Figs. 1a–c demonstrate
that (cid:10)bt(cid:11) tends to vary by more than a factor of two between models. Figs. 1d–f
show an even larger spread for maximal disk models. Figs. 2a-f show this another
way by plotting the rate vs the optical depth for the set of models. These plots
show that (cid:10)bt(cid:11) is indeed much less model dependent than either τ or Γ. While the
rate and τ vary by more than a factor of ten in these plots, their ratio varies only
∼ 2 for a canonical disk. In fact, we note that the line Γ ∝ τ 3/2 is a fairly good
Thus the large scatter in (cid:10)bt(cid:11) seen in
⋆
fit to all the models we have considered.
the maximal disk histograms is mostly just due to the large scatter in rate. (The
⋆ To the extent that the relation Γ = aηm f −1/2τ 3/2 holds, where a is a constant from theory,
m (4/π)3/2 (cid:10)bt(cid:11)−3/2
we have that a = f −1/2η−1
Γ−1/2 is independent of the model parameters.
Thus, if the macho fraction f were known, one could extract the mass integral ηm from
8
rate varies more than the optical depth.) In all the plots we use m = M⊙ , but for
an arbitrary mass distribution just scale Γ by etam and (cid:10)bt(cid:11) by η−1
m . Keep in mind
that a given experiment can produce only one point in the Γ, τ plane and that
the primary use of a measurement will be to find f , the Macho fraction. We see
that for approximate work, the optical depth does a reasonable job of predicting
the rate. But for more detailed work, especially when efficiencies are involved,
the difference between rate and optical depth should be kept in mind. We also
note that the predicted distribution of event durations is found as a differential
rate (Appendix B).
Next let us turn to scatter in the predicted microlensing rate caused by un-
certainties in the halo parameters. Fig. 2 shows that for all lines of sights, there
is a scatter in the rate of more than a factor of ten for a canonical disk. For
the LMC, the models with the smallest rate have spherical halos with small core
radii, falling rotation curves, and small values of v0 , while the models with the
largest rates have either spherical or flattened halos, but large core radii, rising
rotation curves, and large values of v0 . This is as expected, since any model
which puts more mass at a large distance in the direction of the LMC will have
a larger microlensing rate, and a larger optical depth. This is shown in Fig. 3
in which we plot the optical depth against the rotation velocity at r = 50 kpc.
The correlation between τ and v2
c (r = 50 kpc), while not perfect, is quite good.
Note that the mean value of the rotation velocity at R0 is nearly independent of
the microlensing rate. Figs. 2d–f and Fig. 3c-d show the case of a maximal disk.
Here we see that the rate and optical depth can be considerably smaller than
for a canonical disk. Also there is a variation between models of several orders
of magnitude. This is as expected since in these models the disk is the main
contributor to the rotation curve at the solar distance. Thus a smaller enclosed
halo mass is required to match observations, and the halo parameters are poorly
constrained.
The halo may only consist of a fraction f of baryonic matter in the form
of Machos. Thus, a factor of more than ten uncertainty in the predicted rate
caused by the poorly determined halo parameters makes it difficult to determine
the allowed amount of non–baryonic dark matter. It is clearly essential to reduce
the uncertainty.
5. Reducing Model Uncertainties
The primary way of reducing the model uncertainties in the microlensing
observables is to determine the halo parameters. Even within the restricted
framework of the power–law galaxy models, if β , v0 , q , R0 and Rc are known,
there is still uncertainty in the rate. This is because the DFs equation (8) are
the simplest consistent with the potential and the density, but are certainly not
unique. There are still further DFs that depend on non–classical third integrals
of motion and generate anisotropic velocity distributions. Note, too, that even
though our models give a plausible representation of the Milky Way, there cer-
tainly exist other alternatives (see e.g., Frieman & Scoccimarro 1994; Gates &
Turner 1993; Giudice, Mollerach & Roulet 1994) with different lensing properties.
And of course, the size of the disk plays a crucial role.
observables ηm ≈ f 1/2a−1 (4/π)3/2Nef f E 1/2 (P bti/ǫi )−3/2 , where E is the total exposure,
Nef f = P ǫ−1
, ǫi is the efficiency at which events of duration bti are recovered, and the sums
i
go from 1 to the number of observed microlensing events. For LMC microlensing in our set
of models we find a ≈ 3850 ± 260yr−1 . The physical basis for this relationship may simply
be that the optical depth is proportional to the mass along the line of sight ∝ v2
c , and the
rate is proportional to the optical depth times vc .
9
One obvious way to determine halo parameters is to use conventional astro-
nomical techniques – observations of stars, gas and satellites – to fix the solar
radius and circular speed more accurately. For example, fixing the solar radius
at 8 kpc, and demanding vcirc = 220 km/s ±5% between 8 and 16 kpc reduces
the spread in microlensing rates toward the LMC from more than a factor of ten
to a little more than a factor of two (for the canonical disk). Uncertainties in τ
and (cid:10)bt(cid:11) are reduced similarly. A better determination of the halo core radius by
stellar observations would also be important.
However, it is also possible to use the microlensing experiments themselves to
determine the halo parameters and reduce the model uncertainty. The basic idea
is to exploit the fact that there are at least four viable lines-of-sight out of the
Milky Way in which to measure the microlensing rate and average event duration.
Each line-of-sight (LMC, SMC, M31 and the bulge) offers a different “pencil
beam” through the dark halo, and so by comparing the rates, optical depths,
and average durations among the different lines-of-sight information concerning
the halo shape can be gained. Several of the parameters, such as flattening q
and asymptotic slope of the rotation law β , may best be determined this way.
So microlensing gives us a new probe of the density and velocity structure of the
dark halo. This is in addition to information on the size of the disk gained via
microlensing.
For instance, a scatter plot of the ratio of LMC and SMC rates vs the LMC
and SMC average durations is shown in Fig. 4a. The models clearly fall into
two distinct groups. Those models marked with a circle all have round halos
(E0), while those with a square are flattened to roughly E6. Thus the ratio of
LMC rate to SMC rate is a excellent indicator of halo flattening. This effect was
first discovered – using optical depth rather than microlensing rate – by Sackett
& Gould (1993). Frieman & Scoccimarro (1994) have recently cautioned that
the robustness of this diagnostic may be lost if the halo is tilted with respect to
the disc – although such a configuration cannot be a long–lasting equilibrium.
So, the halo flattening can probably be determined if enough events are found
to allow accurate measurement of the SMC microlensing rate. Figs. 4b and 4c
show the rate ratio for M31/LMC, and M31/SMC. While separation of flattened
models is still evident, one sees from the figures that it is the LMC and SMC
position relative to the halo axis of symmetry that make the measurement of
the flattening so easy. Note again, that in an experiment one measures only one
LMC rate (and optical depth) and one SMC rate, and so gets only one point in
any of these scatter plots. It is also interesting to observe from Fig. 4a that the
model uncertainties in the LMC/SMC rate ratio are much greater for flattened
halos than for spherical halos. The case of a maximal disk is not shown, since it
looks almost identical to Fig. 4.
Can we use microlensing to determine whether the halo has a rising or falling
rotation curve? The LMC and SMC are at nearly the same distances (50 and 60
kpc), so it is natural to expect the ratio of M31 to LMC microlensing to be the
most useful discriminant. Note that rate ratios are convenient to use, because the
magnitude of any rate always contains the unknown parameter f . In Fig. 5 we
plot the M31/LMC rate ratio vs the LMC rate for the set of models above, with
triangles for β = 0.2 (falling rotation curve), circles for β = 0 (asymptotically
flat rotation curves), and stars for β = −0.2 (rising rotation curve). In Fig. 5a,
all models are plotted, while in Fig. 5b and Fig. 5c only models with spherical
(q = 1) and flattened (q = 0.71 or q = 0.78) halos respectively are shown.
In Fig. 5a some separation of models with different values of β is evident but
there is substantial ambiguity, which would make a direct estimate of β using
this method difficult. However, suppose that we have already determined the
10
halo flattening by use of the ratio ΓLM C /ΓSM C . Then, as shown in Figs. 5b
and 5c for a canonical disk, a fairly clear separation of rising, falling, and flat
rotation curve parameter can be accomplished. Thus, the ambiguity seen in
Fig. 5a is largely removed when models with different flattenings are plotted
separately. The exception is some overlap between models with Rc = 2 kpc and
Rc = 20 kpc and different values of β . This ambiguity is probably removable
as discussed below. The case of a maximal disk is not displayed, as it is very
similar. So, the asymptotic form of the rotation law, or equivalently β , can
probably be determined from the M31/LMC rate ratio once q is known. Keep in
mind, however, the caveats mentioned in §3 concerning our M31 rate calculation,
which may result in corrections which modify this effect. If halo microlensing
can be distinguished from M31 microlensing, a measurement of β should then be
possible.
Next, can we determine the halo core radius Rc? The parameter Rc affects
mainly the inner portion of the halo and overall normalization of the halo mass.
This overall normalization is mixed in with v0 and f , and so the best hope
in determining Rc is probably a comparison of the bulge with a more distance
source such as the LMC. Here we have the problems mentioned in §3 concerning
bulge microlensing; our modelling of the distribution of velocities is not adequate
along the disk. But, the optical depth is independent of the velocities and will
give some indication of the rate. Even so, our calculations do not give the total
optical depth towards the bulge, merely the contribution of the optical depth
from the halo.
In Fig. 6, we plot the LMC/bulge optical depth ratio vs the bulge optical
depth, where triangles indicate Rc = 2 kpc, boxes indicate Rc = 5 kpc, and
stars indicate Rc = 10 kpc. A reasonably clean separation is obtained when
this ratio is plotted for all the models (not shown). In Fig. 6, this separation
is made clear–cut, if one supposes β and q have already been measured by the
methods above. The Rc = 20 kpc models have an LMC/bulge ratio of greater
than 10, and are very easily distinguished even with no prior knowledge of β
and q . (They fall off the top of the plots in Fig. 6). Even if β and q are not
known, the separation is quite good if the value of the solar radius R0 is held
fixed. So, a better determination of R0 by non-microlensing means can allow a
clearer separation of the effect of the halo core radius. The case of a maximal
disk is not shown since it gives very similar results.
6. Distribution of Event Durations
Since the duration of a microlensing event is proportional to the Einstein
radius (∝ m1/2 ), the duration of an event gives information about the mass of lens
which caused it. In trying to understand the nature of the ob jects responsible for
the observed microlensing, this is important information. But the duration also
depends upon the unknown lens velocity and distance. Thus, a given mass Macho
can cause a wide distribution of event durations. This distribution must be used
statistically to infer probable masses from observed durations. Using the DF’s
(equation (8)), the distribution of event durations can be found. The formula and
definitions are given in Appendix B. In Fig. 7, we show several bt distributions.
One sees that different halo parameters give quite different distributions. It is the
average of these distributions (cid:10)bt(cid:11) that is shown in the histograms in Fig. 1. Fig. 7
shows that, as expected, uncertainty in the halo model will lead to additional
uncertainty in determining the masses of the lensing ob jects. The curves labeled
(a), (b), and (c) are canonical disk cases with various choices of halo parameters,
while curve (d) shows a maximal disk example. We also note that the scaling
11
introduced in Griest (1991) works fairly well for models we considered. That
, and the dΓ/dbt axis by (cid:10)bt(cid:11), all the curves are
is, by scaling the bt axis by (cid:10)bt(cid:11)−1
found to lie roughly on top of each other. This means that for power law galaxy
models along a given line-of-sight, the shape of the distribution is much more
model independent than peak value.
In a future paper we plan to explore further the information that can be
extracted from event duration distributions, and include other possibilities such
triaxiality, streaming motion, etc.
7. Conclusions
This paper has shown how to exploit the power–law galaxy models (E93, E94)
as simple, flexible and realistic representations of the dark halo. These models
have the advantage of simple and analytic phase space distribution functions and
therefore permit accurate calculation of the optical depth, microlensing rate and
average event duration. We provide formulae for these quantities as a function of
the halo parameters and source distance and direction (Appendix A). The distri-
bution of event timescales is presented in Appendix B. We apply our formulae to
study microlensing towards the Large and Small Magellanic Clouds (LMC and
SMC), the galactic bulge, and the M31 disk galaxy. We find that:
(1) For a canonical disk, the optical depth is a reasonable indicator of the mi-
crolensing rate to within a factor of two. This is important, because the optical
depth is much easier to calculate than the rate and probably will continue to be
widely used by investigators. For more accurate work, as well as for derivations
of the distribution of durations, galaxy modeling with distribution functions is
crucial. For a maximal disk the agreement between optical depth and rate is less
robust, though the relation Γ ∝ m−1/2 τ 3/2 seems to hold.
(2) The evaluation of the fraction f of the halo consisting of Machos is hampered
by the uncertainties in the galactic constants, such as the shape of the rotation
law and the flattening of the dark halo. For a realistic set of halo models, we
found rates toward the LMC and SMC can vary by more than a factor of ten
from model to model for a canonical disk, and by several orders of magnitude for
a maximal disk. Left unaddressed, this model uncertainty will thwart accurate
determination of f .
(3) An attractive way of reducing the uncertainty – which simultaneously opens
up a new method in galactic astronomy – is to use microlensing to explore the
shape and structure of the dark halo. This has also been realised by Sackett &
Gould (1993), who showed that the ratios of optical depth towards the LMC and
SMC is a robust indicator of the flattening of the dark halo. We confirm this
result by showing that the ratios of the event rates also distinguish flatness. In
particular, the ratio of microlensing rates towards the LMC and SMC is ∼ 0.7−0.8
for E0 halos and ∼ 1.0 − 1.2 for E7 halos. This is true for both canonical and
maximal disk models. Once the flattening has been established, the asymptotic
slope of the rotation curve β might be determined using the M31/LMC rate ratio.
The LMC/bulge ratio contains important information on the halo core radius.
We caution that this may not be easy to extract, as the dark halo is probably
not the dominant source of lenses towards the bulge.
In summary, the discovery of a dark halo consisting of a significant fraction
of Machos is only the starting point for an exploration of the halo characteristics
which microlensing can help determine.
Acknowledgements
12
KG thanks A.Gould, D.A.Merritt, and D.N.Spergel for help in the early
stages of this pro ject. KG acknowledges a DOE OJI grant, and KG and CWS
thank the Sloan Foundation for their support. Work performed at LLNL is
supported by the DOE under contract W7405-ENG-48. Work performed by the
Center for Particle Astrophysics on the UC campuses is supported in part by the
Office of Science and Technology Centers of NSF under cooperative agreement
AST-8809616. Work performed at MSSSO is supported by the Bilateral Science
and Technology Program of the Australian Department of Industry, Technology
and Regional Development.
Appendix A
In this appendix, we give the formulae for the microlensing rate and optical
depth for the general flattened halo model described in the text (equations 4–7).
The total rate Γ of microlensing in a power–law halo with model parameters
β , v0 , Rc , R0, and q is
I3 .
I1
I2
Γ =
+
+
Γ(nβ )
Γ(dβ )
0 R3β/2
(β + 2)(1 − q2 )
v3
C0uT
c
2cq2√−βL1/2+3β/2
p2πM /M⊙
0 R2+3β/2
v3
Γ(nβ )
C0uT
(β + 2)
c
cq2√−βL5/2+3β/2
p2πM /M⊙
Γ(dβ )
0 R3β/2
(2 − 1+β
v3
q2 )
Γ(nβ − 2/β )
C0uT
c
c√−βL1/2+3β/2
p2πM /M⊙
Γ(dβ − 2/β )
Here, C0 = 1/√GM⊙ , Γ(x) is the gamma function, and the integrals Ii are
dsps(1 − s)(A′s2 + B ′s + C ′ )
1Z0
(D ′s2 + E s′ + F ′)2+3β/2
dsps(1 − s)
1Z0
(D ′s2 + E s′ + F ′ )2+3β/4
dsps(1 − s)
1Z0
(D ′s2 + E s′ + F ′ )1+3β/4
I2 =
I1 =
I3 =
.
(A1)
(A2)
(A3)
with
A′ =3 cos2 b, B ′ = −6R0 cos b cos ℓ/L,
C ′ =2R2
0 cos2 ℓ/L2 + R2
0/L2 + R2
0 sin2 ℓ sin2 b/L2 ,
D ′ = cos2 b + q−2 sin2 b, E ′ = −2R0 cos b cos ℓ/L
F ′ =(R2
c + R2
0 )/L2 .
The quantities b, l are the galactic coordinates of the source star, L is the source
distance, G is Newton’s constant, and c is the speed of light. The constants nβ
13
4
β
+ 2,
if β > 0,
−4
β −
3
2
,
if β < 0,
and dβ have a different form according to whether β is positive or negative
nβ =
dβ =
In the limit β → 0 (the case of an asymptotically flat rotation curve), the ex-
pression for the rate follows from the above by systematic use of the formula
Γ(x + 1/2)/Γ(x) → √x as x → ∞. The optical depth τ is
1Z0
s(1 − s)(A′′s2 + B ′′s + C ′′ )ds
(D ′s2 + E ′s + F ′ )(β+4)/2
0 Rβ
c u2
v2
T
c2 q2Lβ
−4
β − 1,
4
β
+
5
2
,
.
(A6)
if β < 0,
if β > 0,
(A4)
(A5)
τ =
where
A′′ = (1 − β q2 ) cos2 b + (2 − (1 + β )q−2) sin2 b,
B ′′ = −2(1 − β q2 )R0 cos b cos ℓ/L,
0 (1 − β q2 ))/L2 .
C ′′ = (R2
c (1 + 2q2) + R2
(A7)
The quadratures are straightforward to evaluate on the computer.
Appendix B
The distribution of event durations is important for finding the mass of
the lensing ob jects. It is given by the normalized differential microlensing rate
(dΓ/dbt)/Γ, where bt = 2Re/v⊥ , and v⊥ is the speed of the Macho perpendicu-
lar to the line-of-sight. The time the Macho spends inside the Einstein radius,
min )1/2bt, where umin is defined in equation (1), and uT = 1. The
T − u2
te = (u2
4 (cid:10)bt(cid:11).
average duration is related to the average bt by hte i = π
In many cases
it is advantageous to use distributions in bt, since they are independent of the
amplifications.
14
= 8
dΓ
dt
For the model described in the text, we find:
t4 ! (β + 2) β 1+4/β (q−2 − 1) [a1G′ J1 + a2H ′ J2 ]
πc2 L6
uT
R4
C
c t4 (cid:19) β 1+4/β (β + 2)
πc2 (cid:18) L4
uT
a1 J1
q2
R2
πc2 (cid:18) L4
c t4 (cid:19) β 1+2/β (2 − q−2 (1 + β )) a3 J3
uT
R2
+ 8
+ 8
where,
a1 = ( −1 − 4
β , β < 0
1 + 4
β ,
β > 0,
4(β + 4)
,
a2 =
β 2
a3 = ( −1 − 2
β , β < 0
1 + 2
β > 0,
β ,
(B1)
(B2)
,
s2 +
and the integrals Ji are
J1 = Z ds s2(1 − s)2 (cid:12)(cid:12)(cid:12)(cid:12)
s(1 − s)(cid:12)(cid:12)(cid:12)(cid:12)
4/β
K ′
H ′
g1(D ′s2 + E ′s + F ′ )β/2 −
g1
J2 = Z ds s3(1 − s)3 " A′
)# ×
L (cid:19)2
s + (C ′ − 2 (cid:18) R0
B ′
3
3
(cid:12)(cid:12)(cid:12)(cid:12)
s(1 − s)(cid:12)(cid:12)(cid:12)(cid:12)
4/β−1
K ′
H ′
,
g2(D ′s2 + E ′s + F ′ )β/2 −
g2
J3 = Z ds s2(1 − s)2 (cid:12)(cid:12)(cid:12)(cid:12)
s(1 − s)(cid:12)(cid:12)(cid:12)(cid:12)
2/β
H ′
K ′
g3(D ′s2 + E ′s + F ′ )β/2 −
g3
If β < 0, the integrals are evaluated over the interval [0, 1]. If β > 0, then we
must restrict the domain of integration by
(B3)
.
t2 ≥
8βL s(1 − s) (m/M⊙) (D ′s2 + E ′s + F ′ )β/2
(v0cC0 )2 (RC /L)β
.
(B4)
The constants A′ , B ′ , C ′ , D ′ and E ′ are given in Appendix A. The additional
15
constants are
G′ = (cid:18) R0
cos b sin l(cid:19)2
,
L
β (cid:18) RC
L (cid:19)2
L (cid:19)β
K ′ = (cid:18) v0 t
1
β−4 ) (cid:18) L
v0 t (cid:19) 8
β−4
g2 = H ′( β
,
,
m
8
H ′ =
,
L (cC0)2
M⊙
L (cid:19)2
g1 = H ′(−β/4) (cid:18) v0 t
L (cid:19)2
g3 = H ′(−β/2) (cid:18) v0 t
.
,
(B5)
References
Alcock, C. et al., 1993, Nature, 365, 621
Alcock, C. et al., 1994, ApJ, in press
Ashman, K. 1992, PASP, 104, 1109
Aubourg, E. et al., 1993, Nature, 365, 623
Baillon, P., Bouquet, A., Giraud-Heraud, Y., & Kaplan, J. 1993, A&A, 277, 1
Bahcall, J., Schmidt, M., & Soneira, R., 1983, 265, 730
Bahcall, J., 1984, Ap. J., 287, 926.
Binney, J. & Tremaine, S. 1987, Galactic Dynamics (Princeton University Press,
Princeton)
Buchhorn, M. 1992, PhD Thesis, Australin National University
Caldwell, J.A.R., & Ostriker, J.P. 1981, ApJ, 251, 61
Crotts, A.P.S. 1992, ApJ, 399, L43
DeRujula, A., Jetzer, Ph., & Masso, E. 1991, MNRAS, 250, 348
Dubinski, J. & Carlberg, R., 1991, ApJ, 378, 496
Evans, N.W. 1993, MNRAS 260, 191 (E93)
Evans, N.W. 1994, MNRAS, 267, 333 (E94)
Evans, N.W. & Jijina, J. 1994, MNRAS, 267, L21
Fich, M. & Tremaine, S. 1991, ARAA, 29, 409
Fich, M., Blitz, L. & Stark, A.A. 1989, ApJ, 342, 272
Freeman, K.C. 1977, ApJ, 160, 811.
Frieman, J. & Scoccimarro, R. 1994, ApJ, 431, L23.
Gates, E. & Turner, M.S. 1993, preprint FERMILAB-Pub-93/357-A
Giudice, G.F., Mollerach, S., & Roulet, E. 1994, Phys. Rev. D, in press
Gilmore, G, Wyse, R.F.G., & Kuijken, K. 1989, ARAA, 72, 555
Gould, A. 1990, MNRAS, 244, 25
Gould, A. 1992, ApJ, 392, 442
16
Gould, A. 1993a, private communication
Gould, A. 1993b, ApJ, 404, 451
Griest, K. 1991, ApJ, 366, 412
Griest, K. et al. 1991, ApJ, 372, L79
Jetzer, Ph. 1991, in “Atti del Colloquio de Mathematica”, vol 7, ed. CERFIM,
Locarno
Jetzer, Ph. & Masso E., 1994, Phys. Lett. B., 323, 347
Jones, B., et al. 1993, Lick preprint
Katz, N. 1991, ApJ, 368, 325
Kent, S.M. 1992, ApJ, 387, 181
Kerr, F.J. & Lynden-Bell, D. 1986, MNRAS, 221, 1023
Kiraga, M. & Paczy´nski, B., 1994 ApJ, 430, L101
Nemiroff, R.J. 1989, ApJ, 341, 579
Nemiroff, R.J. 1991, A&A, 247, 73
Merrifield, M.R. 1992, AJ, 103, 1552
Oort, J. H. 1960, Bull. Astr. Inst. Netherlands, 6, 249.
Paczy´nski, B. 1986, ApJ, 304, 1
Paczy´nski, B. 1991, ApJ, 371, L63
Primack, J. R., Seckel, D., & Sadoulet, B. 1988, Ann. Rev. Nucl. Part. Sci., 38,
751
Reid, N.J. 1989, in The Center of the Galaxy, IAU Symposium No. 136. ed.
Morris, M. (Kluwer, Dordrecht)
Rohlfs, K., Chini, R., Wink, J.E., & Bohme, R. 1986, A&A, 158, 181
Sackett, P. D. & Gould, A. 1993, ApJ 419, 648
Sahu, K., 1994, Nature, 370, 275
Udalski, A., et al., 1993, Acta Astron, 43, 289
Udalski, A., et al., 1994a, ApJ, 426, L69
Udalski, A., et al., 1994b, Acta Astron, 44, 165
Warren, M.S., et al., 1992, ApJ 339, 405
17
Figure Captions
Figure 1: Histograms of the average duration (cid:10)bt(cid:11) = 4
π τ /Γ for the ensemble of
halo models discussed in the text. Part (a) is for the LMC, (b) is for the SMC,
and (c) is for M31.
If optical depth tracked microlensing rate perfectly each
histogram would be a delta function. Parts (d)-(f ) are the same for a maximal
disk model. Note all plots are for m = 1M⊙ ; scale by η−1
m for other masses
(Equation (3)).
Figure 2: Scatter plots of microlensing rate vs optical depth for the ensemble
of models discussed in the text. Part (a) is for the LMC, (b) is for the SMC, and
(c) is for M31. Each point represent a consistent model of the dark halo. Parts
(d)-(f ) are the same for a maximal disk model. All event rates scale Γ ∝ ηm .
Figure 3: Scatter plots of optical depth vs v2
c (50 kpc), the square of the total
rotation velocity at 50 kpc in the galactic plane. The mass of the Galaxy interior
to this distance is proportional to this squared velocity. Parts (a) and (b) are for
a canonical disk, while parts (c) and (d) are for a maximal disk.
Figure 4: Finding the flattening parameter q . Scatter plots of the ratio of
rates vs. the ratio of event durations. The circles represent halo models which
are spherical (q = 1), while the squares represent flattened halos (q = .71 for
β = 0, −0.1; q = 0.78 for β = 0.1). part (a) is for LMC/SMC and shows clear
separation of spherical and flattened halos. Part (b) is M31/LMC, and part (c)
is M31/SMC.
Figure 5: Finding the asymptotic slope β . Scatter plots of the M31 rate divided
by the LMC rate vs the LMC rate. The stars represent halo models with β = −0.2
(rising rotation curve), the circle models with β = 0 (flat), and the triangles
models with β = 0.2 (falling). Part (a) shows all models, while part (b) shows
only spherical models and part (c) only the flattened models. Separation of the
models becomes easier if the flattening is known. The line of ambiguity in some
panels is due to Rc = 20 kpc models, which can be distinguished as shown in
Figure 6. All event rates scale Γ ∝ ηm .
Figure 6: Finding the core radius Rc . Each panel shows models of definite values
of β (rotation curve slope) and q (flattening). The separation between models is
quite good. The triangles represent Rc = 2 kpc, the squares represent Rc = 5
kpc, and the stars represent Rc = 10 kpc. Panels marked E0 are for spherical
halos, while those marked E6 are for flattened halos. Models with Rc = 20 kpc
were also considered but they are easily distinguished since they typically have
τ (LMC)/τ (bul) > 10 and therefore fall off the top of the figures.
Figure 7: Examples of LMC bt distributions for various model parameters. The
integral under each distribution is unity. The curve marked (a) is for a “standard”
spherical halo (β = 0, q = 1, Rc = 5 kpc, v0 = 200 km/sec, R0 = 8.5 kpc, and
Γ = 1.64 × 10−6 events/yr). Curve (b) has a shorter average duration (β = −0.2,
q = 1, Rc = 5 kpc, v0 = 200 km/sec, R0 = 8.5 kpc, and Γ = 3.9 × 10−6
events/yr). Curve (c) has a longer average duration (β = 0.2, q = 0.78, Rc = 10
kpc, v0 = 210 km/sec, R0 = 8.5 kpc, and Γ = 1.24 × 10−6 events/yr). Finally
18
curve (d) has a maximal disk, which greatly reduces the amount of halo material
(β = 0, q = 1, Rc = 20 kpc, v0 = 90 km/sec, R0 = 7 kpc, and Γ = 9.37 × 10−8
events/yr). The average of each distribution is (cid:10)bt(cid:11). All event rates scale Γ ∝ ηm .
19
This figure "fig1-1.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig2-1.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig1-2.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig2-2.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig1-3.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig2-3.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig1-4.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig2-4.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig1-5.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
This figure "fig2-5.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9411019v1
THEORY OF EXPLORING THE DARK HALO
WITH MICROLENSING : POWER{LAW MODELS
C. Alcock(cid:3);y , R.A. Allsmanz , T.S. Axelrodz , D.P. Bennett(cid:3);y,
K.H. Cook(cid:3);y, N.W. Evans, K.C. Freemanz, K. Griesty;k ,
J. Jijinak , M. Lehnerk , S.L. Marshally;[ , S. Perlmuttery ,
B.A. Petersonz, M.R. Pratty;[, P.J. Quinnz , A.W. Rodgersz,
C.W. Stubbsy;[, W. Sutherland(cid:127)
(The MACHO Collaboration)
(cid:3) Lawrence Livermore National Laboratory, Livermore, CA
y Center for Particle Astrophysics, University of California, Berkeley, CA
z Mt. Stromlo and Siding Spring Observatories,
Australian National University, Weston, ACT , Australia
Theoretical Physics, Department of Physics, University of Oxford, OX NP, UK
k Department of Physics, University of California, San Diego, CA
[ Department of Physics, University of California, Santa Barbara, CA
(cid:127) Astrophysics, Department of Physics, University of Oxford, OX RH, UK
Submitted to the Astrophysical Journal, November,
Abstract
If microlensing of stars by dark matter has been detected (Alcock et al.
; Aubourg et al. ; Udalski et al. ; Alcock et al. ; Udalski et al.
a,b), then the way is open for the development of new methods in galactic
astronomy. This series of papers investigates what microlensing can teach us
about the structure and shape of the dark halo. In this paper we present formulas
for the microlensing rate, optical depth and event duration distributions for a
simple set of axisymmetric disk{halo models. The halos are based on the \power{
law models" (Evans , ) which have simple velocity distributions.
Using these models, we show that there is a large uncertainty in the predicted
microlensing rate because of uncertainty in the halo parameters. For example,
models which reproduce the measured galactic observables to within their errors
still di(cid:11)er in microlensing rate towards the Magellanic Clouds by more than
a factor of ten. We (cid:12)nd that while the more easily computed optical depth
correlates well with microlensing rate, the ratio of optical depth to rate can vary
by a factor of two (or greater if the disk is maximal). Comparison of microlensing
rates towards the Large and Small Magellanic Clouds (LMC and SMC) and
M can be used to aid determinations of the halo (cid:13)attening and rotation curve
slope. For example, the ratio of microlensing rates towards the LMC and SMC is
(cid:24) : (cid:0) : for E halos and (cid:24) : (cid:0) : for E halos (c.f. Sackett & Gould ).
Once the (cid:13)attening has been established, the ratio of microlensing rates towards
M and the LMC may help to distinguish between models with rising, (cid:13)at or
falling rotation curves. Comparison of rates along LMC and galactic bulge lines-
of-sight gives useful information on the halo core radius, although this may not be
so easy to extract in practice. Maximal disk models provide substantially smaller
halo optical depths, shorter event durations and even larger model uncertainties.
Sub ject headings: dark matter - Galaxy: structure - gravitational lensing
. Introduction
The recent detection of possible gravitational microlensing events (Alcock et
al. ; Aubourg et al. ; Udalski et al. , Alcock et al. ; Udalski et al.
a,b) gives hope that at least part of the dark matter content of our galaxy is
directly accessible to observation. The dark halo, whose extent has been studied
gravitationally for many years via velocities of stars, gas, and satellites (e.g. Fich
& Tremaine ), contains at least three times (and perhaps more than ten
times) the mass of the luminous galaxy. Its identity is one of the ma jor unsolved
problems in astronomy (e.g. Ashman ; Primack, Seckel, & Sadoulet ).
While it is possible that the halo consists mostly of exotic non-baryonic elemen-
tary particles, the idea of Paczy(cid:19)nski ( ) of searching for Massive Compact
Halo Ob jects (Machos) in the range (cid:0)M(cid:12) to M(cid:12) by monitoring millions
of stars in the LMC may have borne fruit in the experimental programs.
If the Milky Way halo contains large numbers of Machos, then the gravi-
tational microlensing experiments now under way should have the potential to
determine the number and distribution of Machos in the halo, as well as their
mass distribution. The next step for the microlensing experiments will be to
gather more events and then translate the number and duration of those events
into an estimate of the mass fraction f of the dark halo which consists of Machos
in the relevant mass range. To accomplish this goal a model of the dark halo
is necessary. In the past simple spherical models with (cid:13)at rotation curves have
been considered (Paczy(cid:19)nski ; Griest ; DeRujula, Jetzer & Masso ;
Nemiro(cid:11) ). These have been valuable in estimating the order{of{magnitude
e(cid:11)ects but su(cid:11)er from at least three important de(cid:12)ciencies:
() The halo may not be spherical. N-body simulations of gravitational collapse
of collisionless dark matter generically produce axisymmetric or triaxial halos
(Quinn, et al. ; Dubinski & Carlberg ; Katz ). The recent papers
of Sackett & Gould ( ) and Frieman & Scoccimarro ( ) have made an
important start on the study of microlensing e(cid:11)ects in (cid:13)attened halos (see also
the early work of Jetzer ).
() The e(cid:11)ect of the galactic disk is ignored. The disk makes a signi(cid:12)cant con-
tribution to the local circular speed. Modeling without proper allowance for this
e(cid:11)ect leads us to over{estimate the gravity (cid:12)eld and hence the mass of the dark
halo. That is, since the amount of material in the galactic halo is set by the local
circular speed, a larger contribution to this speed by the disk, means a smaller
halo is needed to explain the total circular velocity.
() It is only a simpli(cid:12)ed view of the data that permits one to regard the rotation
curve of the Milky Way as (cid:13)at. In fact, even the sign of the local gradient of the
rotation law at the sun is not known { Fich, Blitz & Stark ( ) estimate that
it may be rising or falling by about km/sec outwards from the solar circle to
a galactocentric radius of kpc.
In this paper, we make a start towards quantifying and remedying these defects
by calculating the microlensing rate and optical depth in a set of simple, (cid:13)exible
and realistic halo models. We take the power{law models (Evans , ,
hereafter E , E ) { for which simple and self-consistent distribution functions
are known { and provide an approximate method to allow for the in(cid:13)uence of the
galactic disk. This enables calculation not just of the optical depth, but also the
event rate and the distribution of event durations. In this paper, the emphasis is
on the uncertainties in microlensing predictions and what can be done to reduce
them. Since there are many possible models of the dark halo consistent with
current observations, there is a substantial scatter in the predicted microlensing
rate { and this in turn contributes to an uncertainty in the measurement of the
fraction of the halo consisting of Machos. Of course, this is directly relevant to
the di(cid:14)cult but important question of whether the observed microlensing rates
can rule out or support the existence of non{baryonic matter in the halo.
We show that the often calculated optical depth is a useful predictor of the
microlensing rate to a factor of (cid:24) . For more accurate work and to predict event
durations, a distribution of Macho velocities is needed. By considering a large
range of model parameters consistent with observations, we (cid:12)nd the microlensing
optical depth can vary by a factor of six or more.
If the disk of the Milky
Way is \maximal" { in the sense that it provides almost all of the Galacto-
centric acceleration { then a much smaller halo is required. This gives an even
larger spread in predicted optical depth, rate, and event durations. An attractive
possibility is to use the measured microlensing rates toward di(cid:11)erent sources (e.g.
LMC, SMC, M and the galactic bulge) to determine some of the halo and disk
parameters, thereby providing a new tool for the study of galactic structure and
at the same time reducing the halo uncertainty in the measurement of f . We
corroborate the Sackett & Gould ( ) prediction that the ratio of LMC to
SMC optical depth is a robust indicator of the (cid:13)attening of the dark halo { and
extend it by showing that the ratio of microlensing rates distinguishes (cid:13)atness as
well. We predict that the ratio of rates toward M and the LMC may be enable
us to discover whether the rotation curve is rising or falling (or equivalently the
extent of the dark halo). A comparison of the bulge and LMC rates can provide
information on the halo core radius.
The plan of the paper is as follows: In x we review the axisymmetric models
and present formulas for the optical depth, microlensing rate, and event dura-
tions. In x we discuss the halo parameters and their allowed range and explain
how to take into account the e(cid:11)ect of the galactic disk. In x we compare op-
tical depth and rate, and discuss the uncertainties in microlensing rates due to
uncertainties in the parameters. In x we discuss reducing those uncertainties by
comparing results along di(cid:11)erent lines-of-sight, in x we discuss distributions of
event durations, and in x we summarize our conclusions.
. Axisymmetric models
The primary goal of the galactic gravitational microlensing experiments is
to determine the mass of the dark halo in Machos. The experiments search for
Machos by monitoring millions of stars nightly in the Large Magellanic Cloud
(LMC), the Small Magellanic Cloud (SMC), the galactic bulge, and perhaps in the
future in the M galaxy (Crotts ; Baillon et al. ). If the dark halo con-
tains large numbers of Machos, occasionally one passes close to the observer{star
line{of{sight and acts as a gravitational lens, causing a time-dependent magni(cid:12)-
cation of the stellar image. The resulting lightcurve is determined by only a few
quantities such as the distance s from us to the Macho, Macho mass m, Macho
transverse velocity v?, and impact parameter b. The magni(cid:12)cation A(t) as a
function of time t is given by
A(t) = (u + )=[u(u + )= ];
u(t) = b=Re = (umin + ! (t (cid:0) t ) )= ;
()
Re = (=c)[Gms( (cid:0) s=L)]= ;
where the peak magni(cid:12)cation Amax is given by inverting umin = u(Amax ), ! =
v?=Re , and L is the distance to the star. Experimentally, microlensing events
are characterized by the maximum magni(cid:12)cation Amax , the time of the peak t ,
and the duration of the event bt, where bt = =! . Also used in the literature as an
\event duration" is te = bt(uT (cid:0) umin )= . This is time for which A (cid:21) AT , with
AT = A(uT ). Using AT = : corresponds to uT = which is the time inside
the Einstein radius Re . A more thorough discussion can be found in in Paczy(cid:19)nski
( ) and Griest ( ).
An observing team measures the number and duration of microlensing events.
The number of observed events is proportional to the number of stars monitored,
the duration of the experiment, the experimental e(cid:14)ciency, and the rate at which
microlensing occurs. The primary observables are the optical depth (cid:28) , the rate
(cid:0) and the average duration of the events (cid:10)bt(cid:11). They are related by
(cid:28) = (cid:0) hte i = (cid:25) (cid:0) (cid:10)bt(cid:11) :
()
If the distribution of Machos masses n(m) were a delta function, then (cid:0) would
be (cid:0) = (cid:0) (m=M(cid:12) )(cid:0)= , where (cid:0) is the rate with m = M(cid:12) and is independent
of mass. For a general normalized mass distribution (cid:0) = (cid:17)m(cid:0) , where the mass
integral is
(cid:17)m = Z dmn(m)(m=M(cid:12) )(cid:0)= :
()
This also implies that (cid:10)bt(cid:11) = (cid:10)bt(cid:11) (cid:17)(cid:0)m .
The optical depth is the number of Machos inside the microlensing tube with
radius uT Re (s) and length L. It depends only on the density of Machos (cid:26)
(cid:28) = Z (cid:26)(x)m (cid:25)uT Re (s)ds:
()
Unlike the rate or average duration, it is independent of the Macho mass distri-
bution. For this reason it also contains no direct information about the mass of
the lensed ob jects. Note also that since detection e(cid:14)ciencies depend upon the
duration of events, it is important to have models which predict durations. Now
to (cid:12)nd the rate at which Machos enter the microlensing tube requires knowledge
of the distribution of velocities all along the tube. So the rate and the distri-
bution of event durations are hard to calculate because they require the entire
phase space distribution function (DF) F (v; x). The di(cid:11)erential rate is given by:
d(cid:0) = m F (v; x) cos (cid:18)uT Rev?dvdxd(cid:11);
()
where the angles and notation are de(cid:12)ned in Griest ( ).
What makes the calculation particularly di(cid:14)cult is that the DF cannot be
prescribed arbitrarily as a Maxwellian, for instance. This is because the Machos
are collisionless, so the DF is constrained to obey the collisionless Boltzmann
equation. By Jeans' theorem, this implies that the DF depends only on the iso-
lating integrals of motion (see Binney & Tremaine , p. ). Self{consistent
solutions for distributions of velocities that build (cid:13)attened halo models are scarce.
The largest known set of axisymmetric models with simple DFs are the \power{
law galaxies" (E , E ). These form the basis for the exploration of microlensing
in this paper, allowing us to go beyond simple spherical models. Note, however,
that all these models are axisymmetric and oblate, while N-body simulations
suggest that halos may well be triaxial. Exploration of triaxial models will be
done in a future paper.
The parameters of the power{law models are:
() The core radius Rc, which measures the scale at which the density law begins
to soften.
() The (cid:13)attening parameter q , which is the axis ratio of the concentric equipo-
tential spheroids, with q = representing a spherical (E ) halo and q (cid:24) :
representing an ellipticity of about E. The \isophotal" ellipticity of the dark
halo is a function of q , as well as other parameters of the model [see E , eq.
(. )].
() The parameter (cid:12) , which determines whether the rotation curve asymptotically
rises, falls or is (cid:13)at. At large distances R in the equatorial plane, the rotation
velocity vcirc (cid:24) R(cid:0)(cid:12) . So (cid:12) = corresponds to a (cid:13)at rotation curve, while (cid:12) <
is a rising rotation curve and (cid:12) > is falling.
()The solar radius R , which is the distance of the Sun from the galactic center.
() Finally, the normalization velocity v , which determines the overall depth
of the potential well and hence the typical velocities of Machos in the halo. In
the limit (cid:12) = , q = and large R (spherical halo with a (cid:13)at rotation curve)
v = vcirc .
Using z as the height above the equatorial plane, the potential of the power{
law models is
(cid:9) = >>>>>><>>>>>>:
v R(cid:12)c =(cid:12)
(Rc + R + z q(cid:0) )(cid:12) = ;
if (cid:12) = ,
()
(cid:0) v log(Rc + R + z q(cid:0) );
if (cid:12) = ,
and the mass density is
(cid:26) = v R(cid:12)c
(cid:25)Gq Rc ( + q ) + R ( (cid:0) (cid:12) q ) + z ( (cid:0) ( + (cid:12) )q(cid:0) )
:
()
(Rc + R + z q(cid:0) )((cid:12)+)=
The DF corresponding to this potential{density pair is
F (E ; Lz ) = >>>><>>>>: ALz jE j=(cid:12)(cid:0)= + B jE j=(cid:12)(cid:0)= + C jE j=(cid:12)(cid:0)=;
if (cid:12) = ,
ALz exp(E =v ) + B exp(E =v ) + C exp(E =v );
if (cid:12) = .()
where the constants A, B , and C are given in E and E . As required by Jeans'
theorem, the DFs depend only on the isolating integrals of motion, namely the
relative energy per unit mass E = (cid:9) (cid:0) v , and the angular momentum per unit
mass about the symmetry axis Lz . The circular velocity in the equatorial plane
is
v R(cid:12)c R
v circ =
(Rc + R )((cid:12)+)= :
( )
Note that the limit q = , (cid:12) = and Rc = recovers the standard singular
isothermal sphere used by Paczy(cid:19)nski. Allowing a core radius gives
(cid:26) = v (cid:25)G R + Rc
(R + Rc ) :
( )
This di(cid:11)ers from the cored isothermal sphere considered by Griest ( ) in sev-
eral ways. First, the rotation curve approaches its asymptotic value more quickly.
Second, the DF given by the q = , (cid:12) = limit of equation () is self{consistent,
whereas Griest ( ) assumed an approximate Maxwellian distribution of ve-
locities.So far we have only modeled the dark halo. However, in the standard model,
a substantial fraction ((cid:24) %) of the centripetal force at the solar radius derives
from the disk stars. This is represented by a thin exponential disk with a scale
length of Rd = : kpc, normalized to a surface density of (cid:6) = M(cid:12)pc(cid:0) at the
solar radius (Gilmore et al. ; Gould ). It is possible that the disk of our
galaxy is substantially larger than the canonical value. (Oort ; Bahcall ;
Kuijken and Gilmore ; Gould ). Recent microlensing results (Alcock,
et al. ; Udalski, et al. ) as well as studies of the optical rotation curves
of external galaxies (Buchhorn ; Kent ) may suggest this. We consider
such a \maximal disk" by taking (cid:6) = M(cid:12)pc(cid:0) . The rotation velocity added
in quadrature is thus (Freeman ; Binney & Tremaine , p. )
v disk = (cid:25)G(cid:6) hy [I (y )K (y ) (cid:0) I (y )K (y )] ;
()
where y = R=(Rd ), and the In and Kn are modi(cid:12)ed Bessel functions. Note
that in adding a contribution from the disk to the local circular velocity, we have
sacri(cid:12)ced self{consistency. Really, we should (cid:12)nd the DF of the power{law halo
in the combined potential (cid:12)eld of both disk and halo { instead, we use the DF
(). As has been argued elsewhere (Evans & Jijina, ), this is an reasonable
approximation for the LMC, SMC and M, where microlensing typically occurs
at heights above the equatorial plane of many kpc.
In this paper, our aim is to estimate the contribution of the Galactic halo to
microlensing. Of course, this is not the only possible source of de(cid:13)ectors. Towards
the LMC, there is the possibility of microlensing by the LMC dark halo or disk
(Gould b, Sahu ). The optical depth is (cid:24) : (cid:2) (cid:0) for microlensing by
LMC halo lenses and (cid:24) : (cid:2) (cid:0) for LMC disk lenses. The Galactic halo makes
a contribution that is roughly three times greater and so is the dominant source
of lenses. However, this is not the case for lines of sight towards M. The optical
depth is dominated by Machos in the halo and disk of M (Crotts , Gould
a). Crotts ( ) estimates that the halo of our own Galaxy contributes just
% to the total optical depth. Microlensing towards the Galactic bulge poses
perhaps the hardest problems of separating the contributions of di(cid:11)erent de(cid:13)ector
populations. Bulge stars can undergo microlensing not only by halo Machos, but
also by other bulge and disk stars (Griest et al. , Paczy(cid:19)nski , Kiraga
& Paczy(cid:19)nski ). At Baade's Window, the optical depth is (cid:24) : (cid:2) (cid:0) for
microlensing by bulge lenses, (cid:24) : (cid:2) (cid:0) for disk lenses. The dark halo only
makes an important contribution if the core radius is small.
We are now in a position to calculate the microlensing observables { the
optical depth, rate and average duration of events. They can be found using
equations , , and . The results are single quadratures and readily evaluated
on the computer. They are displayed in Appendix A. In Appendix B, we give
the di(cid:11)erential microlensing rate d(cid:0)=dbt, where bt is de(cid:12)ned just after equation ().
The probability of obtaining an event of duration bt is just (d(cid:0)=dbt)=(cid:0).
. Range of models
In order to explore the scatter in microlensing observables, we build a set
of halo models which span the observationally allowed range. The power{law
galaxy models allow us to vary the (cid:13)attening, core radius and rotation law, and
we consider both canonical and maximal disks. For each parameter in the model,
we therefore (cid:12)nd the range permitted by the observations. Then, several values
of each parameter are chosen to represent the range. We also ensure that each set
of parameters gives a model consistent with the measured Milky Way rotation
curve. So, we study the statistical properties of an ensemble of models, each one
of which is a plausible representation of the dark halo of the Milky Way.
For the dark halo (cid:13)attening, little is known. So the entire range of (cid:13)atten-
ing allowed by the power{law models is examined. This varies between E or
spherical (q = ) and roughly E or E (depending on (cid:12) ). The core radius of the
dark halo is also uncertain { Bahcall, Schmidt & Soneria ( ) estimate Rc as
kpc from star count data, while Caldwell & Ostriker ( ) suggest kpc. If
the disk is maximal, values as large as kpc are possible. We consider values
of kpc, kpc, kpc, and kpc. The parameter (cid:12) determines the slope of
asymptotic circular velocity. Between R and R , the circular velocity is prob-
ably within (cid:0) % of the I.A.U value of km/s, but whether the measured
HI rotational velocities rise or fall with R depends upon estimates of the solar
position R and the local circular speed vcirc (R ) (see Fich et al. ; Jones et
al. ). Beyond kpc, little is known directly, though arguments based on
the kinematics of distant satellite galaxies support the idea of a relatively (cid:13)at
rotation curve out to an unknown cut{o(cid:11) (Fich & Tremaine ). However,
current theories of galaxy formation tend to favor the alternative view that dark
halos extend inde(cid:12)nitely, fading into structure on larger scales. So, we do not
consider a halo cut{o(cid:11) in this paper { it would add yet another poorly known
parameter to our model. We investigate power{law halos with (cid:12) = - ., , and
.. These correspond to rotation curves which rise by (cid:24)%, are (cid:13)at, or fall
by (cid:24)% between the solar radius and twice the solar radius, depending a little
upon Rc.The value of the solar radius R has been reviewed by Reid ( ). He
shows that most recent determinations lie between kpc and kpc, with .
kpc being his preferred value. This di(cid:11)ers considerably from the IAU value of
. kpc (Kerr & Lynden{Bell ). We examine the values R = , , and
kpc. Finally, perhaps the single most important parameter is the normalization
velocity v . Given our (cid:12)xed disk contribution to the total rotation law, the
parameter v is now speci(cid:12)ed once we settle upon a choice for vcirc (R ). Merri(cid:12)eld
( ) estimates vcirc (R ) = (cid:6) km/s, Fich, Blitz, & Stark ( ) give
vcirc (R ) = (cid:6) km/s, while Rohlfs et al. ( ) give values between
km/s and km/s between R = kpc and R = kpc. For our ensemble of
models, we impose the constraint that the total circular velocity lies between
km/s and km/s at R and R . Note that the IAU value is km/s (Kerr
& Lynden-Bell ). We also investigated a more restricted ensemble of models
with (cid:20) vcirc (R ) (cid:20) km/s. We (cid:12)nd all our results also hold for this more
restricted ensemble.
. Uncertainties in the Rates
First, let us consider the di(cid:11)erence caused by using the optical depth instead
of the microlensing event rate. The optical depth to microlensing is the mean
number of Machos in the microlensing tube; that is the number of microlensing
events taking place at a given moment. It is easy to calculate since it is inde-
pendent of lens mass and velocity, and only requires knowledge of the density
distribution (cid:26)(x). For this reason, it is the most widely estimated quantity. But
how well does it trace the microlensing rate?
We are able to answer this question since both the rate (cid:0) (equation A) and
the optical depth (cid:28) (equation A) are known for the power{law models. One way
to test this is to plot (cid:10)bt(cid:11), which is the ratio of optical depth (cid:28) and (cid:0), (cid:10)bt(cid:11) = (cid:25) (cid:28) =(cid:0),
for many di(cid:11)erent models. The average duration (cid:10)bt(cid:11) is a constant if (cid:28) and (cid:0) are
well{correlated. In Fig. , we show histograms of (cid:10)bt(cid:11) for microlensing towards
the LMC, SMC and M for our ensemble of models. Figs. a{c demonstrate
that (cid:10)bt(cid:11) tends to vary by more than a factor of two between models. Figs. d{f
show an even larger spread for maximal disk models. Figs. a-f show this another
way by plotting the rate vs the optical depth for the set of models. These plots
show that (cid:10)bt(cid:11) is indeed much less model dependent than either (cid:28) or (cid:0). While the
rate and (cid:28) vary by more than a factor of ten in these plots, their ratio varies only
(cid:24) for a canonical disk. In fact, we note that the line (cid:0) / (cid:28) = is a fairly good
(cid:12)t to all the models we have considered. ? Thus the large scatter in (cid:10)bt(cid:11) seen in
the maximal disk histograms is mostly just due to the large scatter in rate. (The
? To the extent that the relation (cid:0) = a(cid:17)mf (cid:0)= (cid:28) = holds, where a is a constant from theory,
we have that a = f (cid:0)=(cid:17)(cid:0)m (=(cid:25))= (cid:10)bt(cid:11)(cid:0)= (cid:0)(cid:0)= is independent of the model parameters.
Thus, if the macho fraction f were known, one could extract the mass integral (cid:17)m from
rate varies more than the optical depth.) In all the plots we use m = M(cid:12) , but for
an arbitrary mass distribution just scale (cid:0) by etam and (cid:10)bt(cid:11) by (cid:17)(cid:0)m . Keep in mind
that a given experiment can produce only one point in the (cid:0), (cid:28) plane and that
the primary use of a measurement will be to (cid:12)nd f , the Macho fraction. We see
that for approximate work, the optical depth does a reasonable job of predicting
the rate. But for more detailed work, especially when e(cid:14)ciencies are involved,
the di(cid:11)erence between rate and optical depth should be kept in mind. We also
note that the predicted distribution of event durations is found as a di(cid:11)erential
rate (Appendix B).
Next let us turn to scatter in the predicted microlensing rate caused by un-
certainties in the halo parameters. Fig. shows that for all lines of sights, there
is a scatter in the rate of more than a factor of ten for a canonical disk. For
the LMC, the models with the smallest rate have spherical halos with small core
radii, falling rotation curves, and small values of v , while the models with the
largest rates have either spherical or (cid:13)attened halos, but large core radii, rising
rotation curves, and large values of v . This is as expected, since any model
which puts more mass at a large distance in the direction of the LMC will have
a larger microlensing rate, and a larger optical depth. This is shown in Fig.
in which we plot the optical depth against the rotation velocity at r = kpc.
The correlation between (cid:28) and v c (r = kpc), while not perfect, is quite good.
Note that the mean value of the rotation velocity at R is nearly independent of
the microlensing rate. Figs. d{f and Fig. c-d show the case of a maximal disk.
Here we see that the rate and optical depth can be considerably smaller than
for a canonical disk. Also there is a variation between models of several orders
of magnitude. This is as expected since in these models the disk is the main
contributor to the rotation curve at the solar distance. Thus a smaller enclosed
halo mass is required to match observations, and the halo parameters are poorly
constrained.
The halo may only consist of a fraction f of baryonic matter in the form
of Machos. Thus, a factor of more than ten uncertainty in the predicted rate
caused by the poorly determined halo parameters makes it di(cid:14)cult to determine
the allowed amount of non{baryonic dark matter. It is clearly essential to reduce
the uncertainty.
. Reducing Model Uncertainties
The primary way of reducing the model uncertainties in the microlensing
observables is to determine the halo parameters. Even within the restricted
framework of the power{law galaxy models, if (cid:12) , v , q , R and Rc are known,
there is still uncertainty in the rate. This is because the DFs equation () are
the simplest consistent with the potential and the density, but are certainly not
unique. There are still further DFs that depend on non{classical third integrals
of motion and generate anisotropic velocity distributions. Note, too, that even
though our models give a plausible representation of the Milky Way, there cer-
tainly exist other alternatives (see e.g., Frieman & Scoccimarro ; Gates &
Turner ; Giudice, Mollerach & Roulet ) with di(cid:11)erent lensing properties.
And of course, the size of the disk plays a crucial role.
observables (cid:17)m (cid:25) f =a(cid:0) (=(cid:25))=Nef f E =(P bti=(cid:15)i)(cid:0)=, where E is the total exposure,
Nef f = P (cid:15)(cid:0)i
, (cid:15)i is the e(cid:14)ciency at which events of duration bti are recovered, and the sums
go from to the number of observed microlensing events. For LMC microlensing in our set
of models we (cid:12)nd a (cid:25) (cid:6) yr(cid:0) . The physical basis for this relationship may simply
be that the optical depth is proportional to the mass along the line of sight / vc , and the
rate is proportional to the optical depth times vc .
One obvious way to determine halo parameters is to use conventional astro-
nomical techniques { observations of stars, gas and satellites { to (cid:12)x the solar
radius and circular speed more accurately. For example, (cid:12)xing the solar radius
at kpc, and demanding vcirc = km/s (cid:6)% between and kpc reduces
the spread in microlensing rates toward the LMC from more than a factor of ten
to a little more than a factor of two (for the canonical disk). Uncertainties in (cid:28)
and (cid:10)bt(cid:11) are reduced similarly. A better determination of the halo core radius by
stellar observations would also be important.
However, it is also possible to use the microlensing experiments themselves to
determine the halo parameters and reduce the model uncertainty. The basic idea
is to exploit the fact that there are at least four viable lines-of-sight out of the
Milky Way in which to measure the microlensing rate and average event duration.
Each line-of-sight (LMC, SMC, M and the bulge) o(cid:11)ers a di(cid:11)erent \pencil
beam" through the dark halo, and so by comparing the rates, optical depths,
and average durations among the di(cid:11)erent lines-of-sight information concerning
the halo shape can be gained. Several of the parameters, such as (cid:13)attening q
and asymptotic slope of the rotation law (cid:12) , may best be determined this way.
So microlensing gives us a new probe of the density and velocity structure of the
dark halo. This is in addition to information on the size of the disk gained via
microlensing.
For instance, a scatter plot of the ratio of LMC and SMC rates vs the LMC
and SMC average durations is shown in Fig. a. The models clearly fall into
two distinct groups. Those models marked with a circle all have round halos
(E ), while those with a square are (cid:13)attened to roughly E. Thus the ratio of
LMC rate to SMC rate is a excellent indicator of halo (cid:13)attening. This e(cid:11)ect was
(cid:12)rst discovered { using optical depth rather than microlensing rate { by Sackett
& Gould ( ). Frieman & Scoccimarro ( ) have recently cautioned that
the robustness of this diagnostic may be lost if the halo is tilted with respect to
the disc { although such a con(cid:12)guration cannot be a long{lasting equilibrium.
So, the halo (cid:13)attening can probably be determined if enough events are found
to allow accurate measurement of the SMC microlensing rate. Figs. b and c
show the rate ratio for M/LMC, and M/SMC. While separation of (cid:13)attened
models is still evident, one sees from the (cid:12)gures that it is the LMC and SMC
position relative to the halo axis of symmetry that make the measurement of
the (cid:13)attening so easy. Note again, that in an experiment one measures only one
LMC rate (and optical depth) and one SMC rate, and so gets only one point in
any of these scatter plots. It is also interesting to observe from Fig. a that the
model uncertainties in the LMC/SMC rate ratio are much greater for (cid:13)attened
halos than for spherical halos. The case of a maximal disk is not shown, since it
looks almost identical to Fig. .
Can we use microlensing to determine whether the halo has a rising or falling
rotation curve? The LMC and SMC are at nearly the same distances ( and
kpc), so it is natural to expect the ratio of M to LMC microlensing to be the
most useful discriminant. Note that rate ratios are convenient to use, because the
magnitude of any rate always contains the unknown parameter f . In Fig. we
plot the M/LMC rate ratio vs the LMC rate for the set of models above, with
triangles for (cid:12) = : (falling rotation curve), circles for (cid:12) = (asymptotically
(cid:13)at rotation curves), and stars for (cid:12) = (cid:0) : (rising rotation curve). In Fig. a,
all models are plotted, while in Fig. b and Fig. c only models with spherical
(q = ) and (cid:13)attened (q = : or q = :) halos respectively are shown.
In Fig. a some separation of models with di(cid:11)erent values of (cid:12) is evident but
there is substantial ambiguity, which would make a direct estimate of (cid:12) using
this method di(cid:14)cult. However, suppose that we have already determined the
halo (cid:13)attening by use of the ratio (cid:0)LM C =(cid:0)SM C . Then, as shown in Figs. b
and c for a canonical disk, a fairly clear separation of rising, falling, and (cid:13)at
rotation curve parameter can be accomplished. Thus, the ambiguity seen in
Fig. a is largely removed when models with di(cid:11)erent (cid:13)attenings are plotted
separately. The exception is some overlap between models with Rc = kpc and
Rc = kpc and di(cid:11)erent values of (cid:12) . This ambiguity is probably removable
as discussed below. The case of a maximal disk is not displayed, as it is very
similar. So, the asymptotic form of the rotation law, or equivalently (cid:12) , can
probably be determined from the M/LMC rate ratio once q is known. Keep in
mind, however, the caveats mentioned in x concerning our M rate calculation,
which may result in corrections which modify this e(cid:11)ect.
If halo microlensing
can be distinguished from M microlensing, a measurement of (cid:12) should then be
possible.Next, can we determine the halo core radius Rc? The parameter Rc a(cid:11)ects
mainly the inner portion of the halo and overall normalization of the halo mass.
This overall normalization is mixed in with v and f , and so the best hope
in determining Rc is probably a comparison of the bulge with a more distance
source such as the LMC. Here we have the problems mentioned in x concerning
bulge microlensing; our modelling of the distribution of velocities is not adequate
along the disk. But, the optical depth is independent of the velocities and will
give some indication of the rate. Even so, our calculations do not give the total
optical depth towards the bulge, merely the contribution of the optical depth
from the halo.
In Fig. , we plot the LMC/bulge optical depth ratio vs the bulge optical
depth, where triangles indicate Rc = kpc, boxes indicate Rc = kpc, and
stars indicate Rc = kpc. A reasonably clean separation is obtained when
this ratio is plotted for all the models (not shown). In Fig. , this separation
is made clear{cut, if one supposes (cid:12) and q have already been measured by the
methods above. The Rc = kpc models have an LMC/bulge ratio of greater
than , and are very easily distinguished even with no prior knowledge of (cid:12)
and q . (They fall o(cid:11) the top of the plots in Fig. ). Even if (cid:12) and q are not
known, the separation is quite good if the value of the solar radius R is held
(cid:12)xed. So, a better determination of R by non-microlensing means can allow a
clearer separation of the e(cid:11)ect of the halo core radius. The case of a maximal
disk is not shown since it gives very similar results.
. Distribution of Event Durations
Since the duration of a microlensing event is proportional to the Einstein
radius (/ m= ), the duration of an event gives information about the mass of lens
which caused it. In trying to understand the nature of the ob jects responsible for
the observed microlensing, this is important information. But the duration also
depends upon the unknown lens velocity and distance. Thus, a given mass Macho
can cause a wide distribution of event durations. This distribution must be used
statistically to infer probable masses from observed durations. Using the DF's
(equation ()), the distribution of event durations can be found. The formula and
de(cid:12)nitions are given in Appendix B. In Fig. , we show several bt distributions.
One sees that di(cid:11)erent halo parameters give quite di(cid:11)erent distributions. It is the
average of these distributions (cid:10)bt(cid:11) that is shown in the histograms in Fig. . Fig.
shows that, as expected, uncertainty in the halo model will lead to additional
uncertainty in determining the masses of the lensing ob jects. The curves labeled
(a), (b), and (c) are canonical disk cases with various choices of halo parameters,
while curve (d) shows a maximal disk example. We also note that the scaling
introduced in Griest ( ) works fairly well for models we considered. That
is, by scaling the bt axis by (cid:10)bt(cid:11)(cid:0) , and the d(cid:0)=dbt axis by (cid:10)bt(cid:11), all the curves are
found to lie roughly on top of each other. This means that for power law galaxy
models along a given line-of-sight, the shape of the distribution is much more
model independent than peak value.
In a future paper we plan to explore further the information that can be
extracted from event duration distributions, and include other possibilities such
triaxiality, streaming motion, etc.. Conclusions
This paper has shown how to exploit the power{law galaxy models (E , E )
as simple, (cid:13)exible and realistic representations of the dark halo. These models
have the advantage of simple and analytic phase space distribution functions and
therefore permit accurate calculation of the optical depth, microlensing rate and
average event duration. We provide formulae for these quantities as a function of
the halo parameters and source distance and direction (Appendix A). The distri-
bution of event timescales is presented in Appendix B. We apply our formulae to
study microlensing towards the Large and Small Magellanic Clouds (LMC and
SMC), the galactic bulge, and the M disk galaxy. We (cid:12)nd that:
() For a canonical disk, the optical depth is a reasonable indicator of the mi-
crolensing rate to within a factor of two. This is important, because the optical
depth is much easier to calculate than the rate and probably will continue to be
widely used by investigators. For more accurate work, as well as for derivations
of the distribution of durations, galaxy modeling with distribution functions is
crucial. For a maximal disk the agreement between optical depth and rate is less
robust, though the relation (cid:0) / m(cid:0)=(cid:28) = seems to hold.
() The evaluation of the fraction f of the halo consisting of Machos is hampered
by the uncertainties in the galactic constants, such as the shape of the rotation
law and the (cid:13)attening of the dark halo. For a realistic set of halo models, we
found rates toward the LMC and SMC can vary by more than a factor of ten
from model to model for a canonical disk, and by several orders of magnitude for
a maximal disk. Left unaddressed, this model uncertainty will thwart accurate
determination of f .
() An attractive way of reducing the uncertainty { which simultaneously opens
up a new method in galactic astronomy { is to use microlensing to explore the
shape and structure of the dark halo. This has also been realised by Sackett &
Gould ( ), who showed that the ratios of optical depth towards the LMC and
SMC is a robust indicator of the (cid:13)attening of the dark halo. We con(cid:12)rm this
result by showing that the ratios of the event rates also distinguish (cid:13)atness. In
particular, the ratio of microlensing rates towards the LMC and SMC is (cid:24) :(cid:0) :
for E halos and (cid:24) : (cid:0) : for E halos. This is true for both canonical and
maximal disk models. Once the (cid:13)attening has been established, the asymptotic
slope of the rotation curve (cid:12) might be determined using the M/LMC rate ratio.
The LMC/bulge ratio contains important information on the halo core radius.
We caution that this may not be easy to extract, as the dark halo is probably
not the dominant source of lenses towards the bulge.
In summary, the discovery of a dark halo consisting of a signi(cid:12)cant fraction
of Machos is only the starting point for an exploration of the halo characteristics
which microlensing can help determine.
Acknowledgements
KG thanks A.Gould, D.A.Merritt, and D.N.Spergel for help in the early
stages of this pro ject. KG acknowledges a DOE OJI grant, and KG and CWS
thank the Sloan Foundation for their support. Work performed at LLNL is
supported by the DOE under contract W -ENG-. Work performed by the
Center for Particle Astrophysics on the UC campuses is supported in part by the
O(cid:14)ce of Science and Technology Centers of NSF under cooperative agreement
AST- . Work performed at MSSSO is supported by the Bilateral Science
and Technology Program of the Australian Department of Industry, Technology
and Regional Development.
Appendix A
In this appendix, we give the formulae for the microlensing rate and optical
depth for the general (cid:13)attened halo model described in the text (equations {).
The total rate (cid:0) of microlensing in a power{law halo with model parameters
(cid:12) ; v ; Rc ; R , and q is
p(cid:25)M =M(cid:12) v R(cid:12) =
((cid:12) + )( (cid:0) q )
cq p(cid:0)(cid:12)L=+(cid:12) = (cid:0)(n(cid:12) )
(cid:0) = C uT
c
(cid:0)(d(cid:12) ) I
p(cid:25)M =M(cid:12) v R+(cid:12) =
cq p(cid:0)(cid:12)L=+(cid:12) = (cid:0)(n(cid:12) )
+ C uT
((cid:12) + )
c
(cid:0)(d(cid:12) ) I
(A)
p(cid:25)M =M(cid:12) v R(cid:12) =
( (cid:0) +(cid:12)q )
c
cp(cid:0)(cid:12)L=+(cid:12) = (cid:0)(n(cid:12) (cid:0) =j(cid:12) j)
+ C uT
(cid:0)(d(cid:12) (cid:0) =j(cid:12) j) I :
Here, C = =pGM(cid:12) , (cid:0)(x) is the gamma function, and the integrals Ii are
I = Z dsps( (cid:0) s)(A s + B s + C )
(D s + E s + F )+(cid:12) =
I = Z
dsps( (cid:0) s)
(A)
(D s + E s + F )+(cid:12) =
dsps( (cid:0) s)
I = Z
(D s + E s + F )+(cid:12) = :
with
A = cos b; B = (cid:0)R cos b cos `=L;
C =R =L + R cos `=L + R sin ` sin b=L ;
(A)
D = cos b + q(cid:0) sin b; E = (cid:0)R cos b cos `=L
F =(Rc + R )=L :
The quantities b; l are the galactic coordinates of the source star, L is the source
distance, G is Newton's constant, and c is the speed of light. The constants n(cid:12)
and d(cid:12) have a di(cid:11)erent form according to whether (cid:12) is positive or negative
n(cid:12) = >>><>>>: (cid:0)(cid:12) (cid:0) ;
if (cid:12) < ,
(A)
(cid:12) + ;
if (cid:12) > ,
d(cid:12) = >>><>>>: (cid:0)(cid:12) (cid:0) ;
if (cid:12) < ,
(A)
(cid:12) + ;
if (cid:12) > ,
In the limit (cid:12) ! (the case of an asymptotically (cid:13)at rotation curve), the ex-
pression for the rate follows from the above by systematic use of the formula
(cid:0)(x + =)=(cid:0)(x) ! px as x ! . The optical depth (cid:28) is
Z s( (cid:0) s)(A s + B s + C )ds
(cid:28) = v R(cid:12)c uT
(A)
:
cq L(cid:12)
(D s + E s + F )((cid:12)+)=
where
A = ( (cid:0) (cid:12) q ) cos b + ( (cid:0) ( + (cid:12) )q(cid:0) ) sin b;
B = (cid:0)( (cid:0) (cid:12) q )R cos b cos `=L;
(A)
C = (Rc ( + q ) + R ( (cid:0) (cid:12) q ))=L :
The quadratures are straightforward to evaluate on the computer.
Appendix B
The distribution of event durations is important for (cid:12)nding the mass of
the lensing ob jects. It is given by the normalized di(cid:11)erential microlensing rate
(d(cid:0)=dbt)=(cid:0), where bt = Re=v?, and v? is the speed of the Macho perpendicu-
lar to the line-of-sight. The time the Macho spends inside the Einstein radius,
te = (uT (cid:0) umin)= bt, where umin is de(cid:12)ned in equation (), and uT = . The
average duration is related to the average bt by hte i = (cid:25) (cid:10)bt(cid:11).
In many cases
it is advantageous to use distributions in bt, since they are independent of the
ampli(cid:12)cations.
For the model described in the text, we (cid:12)nd:
d(cid:0)d^t = uT(cid:25)c LRC ^t ! ((cid:12) + ) j(cid:12) j+=(cid:12) (q(cid:0) (cid:0) ) [aG J + aH J ]
+ uT(cid:25)c (cid:18) LRc ^t (cid:19) j(cid:12) j+=(cid:12) ((cid:12) + )
(B)
a J
q
+ uT(cid:25)c (cid:18) LRc ^t (cid:19) j(cid:12) j+=(cid:12) ( (cid:0) q(cid:0) ( + (cid:12) )) a J
where,
a = ( (cid:0) (cid:0) (cid:12) ; (cid:12) <
+ (cid:12) ;
(cid:12) > ,
a = ((cid:12) + )
(B)
;
(cid:12)
a = ( (cid:0) (cid:0) (cid:12) ; (cid:12) <
+ (cid:12) ;
(cid:12) > ,
and the integrals Ji are
g (D s + E s + F )(cid:12) = (cid:0) H g s( (cid:0) s)(cid:12)(cid:12)(cid:12)(cid:12)=(cid:12) ;
J = Z ds s ( (cid:0) s) (cid:12)(cid:12)(cid:12)(cid:12)
K
J = Z ds s ( (cid:0) s) " A s + B s + (C (cid:0) (cid:18) R L (cid:19) )# (cid:2)
(B)
(cid:12)(cid:12)(cid:12)(cid:12)
g (D s + E s + F )(cid:12) = (cid:0) H g s( (cid:0) s)(cid:12)(cid:12)(cid:12)(cid:12)=(cid:12)(cid:0) ;
K
J = Z ds s ( (cid:0) s) (cid:12)(cid:12)(cid:12)(cid:12)
g (D s + E s + F )(cid:12) = (cid:0) H g s( (cid:0) s)(cid:12)(cid:12)(cid:12)(cid:12)=(cid:12) :
K
If (cid:12) < , the integrals are evaluated over the interval [ ; ]. If (cid:12) > , then we
must restrict the domain of integration by
^t (cid:21) (cid:12)L s( (cid:0) s) (m=M(cid:12) ) (D s + E s + F )(cid:12) =
:
(B)
(v cC ) (RC =L)(cid:12)
The constants A , B , C , D and E are given in Appendix A. The additional
constants areG = (cid:18) R L cos b sin l(cid:19) ;
L (cC ) mM(cid:12) ;
H =
K = (cid:18) v ^tL (cid:19) (cid:12) (cid:18) RCL (cid:19)(cid:12) ;
g = H ((cid:0)(cid:12) =) (cid:18) v ^tL (cid:19) ;
(B)
g = H ( (cid:12)(cid:12)(cid:0) ) (cid:18) Lv ^t (cid:19) (cid:12)(cid:0) ;
g = H ((cid:0)(cid:12) =) (cid:18) v ^tL (cid:19) :
References
Alcock, C. et al., , Nature, ,
Alcock, C. et al., , ApJ, in press
Ashman, K. , PASP, ,
Aubourg, E. et al., , Nature, ,
Baillon, P., Bouquet, A., Giraud-Heraud, Y., & Kaplan, J. , A&A, ,
Bahcall, J., Schmidt, M., & Soneira, R., , ,
Bahcall, J., , Ap. J., , .
Binney, J. & Tremaine, S. , Galactic Dynamics (Princeton University Press,
Princeton)
Buchhorn, M. , PhD Thesis, Australin National University
Caldwell, J.A.R., & Ostriker, J.P. , ApJ, ,
Crotts, A.P.S. , ApJ, , L
DeRujula, A., Jetzer, Ph., & Masso, E. , MNRAS, ,
Dubinski, J. & Carlberg, R., , ApJ, ,
Evans, N.W. , MNRAS , (E )
Evans, N.W. , MNRAS, , (E )
Evans, N.W. & Jijina, J. , MNRAS, , L
Fich, M. & Tremaine, S. , ARAA, ,
Fich, M., Blitz, L. & Stark, A.A. , ApJ, ,
Freeman, K.C. , ApJ, , .
Frieman, J. & Scoccimarro, R. , ApJ, , L.
Gates, E. & Turner, M.S. , preprint FERMILAB-Pub- /-A
Giudice, G.F., Mollerach, S., & Roulet, E. , Phys. Rev. D, in press
Gilmore, G, Wyse, R.F.G., & Kuijken, K. , ARAA, ,
Gould, A. , MNRAS, ,
Gould, A. , ApJ, ,
Gould, A. a, private communication
Gould, A. b, ApJ, ,
Griest, K. , ApJ, ,
Griest, K. et al. , ApJ, , L
Jetzer, Ph. , in \Atti del Colloquio de Mathematica", vol , ed. CERFIM,
Locarno
Jetzer, Ph. & Masso E., , Phys. Lett. B., ,
Jones, B., et al. , Lick preprint
Katz, N. , ApJ, ,
Kent, S.M. , ApJ, ,
Kerr, F.J. & Lynden-Bell, D. , MNRAS, ,
Kiraga, M. & Paczy(cid:19)nski, B., ApJ, , L
Nemiro(cid:11), R.J. , ApJ, ,
Nemiro(cid:11), R.J. , A&A, ,
Merri(cid:12)eld, M.R. , AJ, ,
Oort, J. H. , Bull. Astr. Inst. Netherlands, , .
Paczy(cid:19)nski, B. , ApJ, ,
Paczy(cid:19)nski, B. , ApJ, , L
Primack, J. R., Seckel, D., & Sadoulet, B. , Ann. Rev. Nucl. Part. Sci., ,
Reid, N.J. , in The Center of the Galaxy, IAU Symposium No. . ed.
Morris, M. (Kluwer, Dordrecht)
Rohlfs, K., Chini, R., Wink, J.E., & Bohme, R. , A&A, ,
Sackett, P. D. & Gould, A. , ApJ ,
Sahu, K., , Nature, ,
Udalski, A., et al., , Acta Astron, ,
Udalski, A., et al., a, ApJ, , L
Udalski, A., et al., b, Acta Astron, ,
Warren, M.S., et al., , ApJ ,
Figure Captions
Figure : Histograms of the average duration (cid:10)bt(cid:11) = (cid:25) (cid:28) =(cid:0) for the ensemble of
halo models discussed in the text. Part (a) is for the LMC, (b) is for the SMC,
If optical depth tracked microlensing rate perfectly each
and (c) is for M.
histogram would be a delta function. Parts (d)-(f ) are the same for a maximal
disk model. Note all plots are for m = M(cid:12) ; scale by (cid:17)(cid:0)m for other masses
(Equation ()).
Figure : Scatter plots of microlensing rate vs optical depth for the ensemble
of models discussed in the text. Part (a) is for the LMC, (b) is for the SMC, and
(c) is for M. Each point represent a consistent model of the dark halo. Parts
(d)-(f ) are the same for a maximal disk model. All event rates scale (cid:0) / (cid:17)m.
Figure : Scatter plots of optical depth vs v c ( kpc), the square of the total
rotation velocity at kpc in the galactic plane. The mass of the Galaxy interior
to this distance is proportional to this squared velocity. Parts (a) and (b) are for
a canonical disk, while parts (c) and (d) are for a maximal disk.
Figure : Finding the (cid:13)attening parameter q . Scatter plots of the ratio of
rates vs. the ratio of event durations. The circles represent halo models which
are spherical (q = ), while the squares represent (cid:13)attened halos (q = : for
(cid:12) = ; (cid:0) :; q = : for (cid:12) = :). part (a) is for LMC/SMC and shows clear
separation of spherical and (cid:13)attened halos. Part (b) is M/LMC, and part (c)
is M/SMC.
Figure : Finding the asymptotic slope (cid:12) . Scatter plots of the M rate divided
by the LMC rate vs the LMC rate. The stars represent halo models with (cid:12) = (cid:0) :
(rising rotation curve), the circle models with (cid:12) = ((cid:13)at), and the triangles
models with (cid:12) = : (falling). Part (a) shows all models, while part (b) shows
only spherical models and part (c) only the (cid:13)attened models. Separation of the
models becomes easier if the (cid:13)attening is known. The line of ambiguity in some
panels is due to Rc = kpc models, which can be distinguished as shown in
Figure . All event rates scale (cid:0) / (cid:17)m .
Figure : Finding the core radius Rc . Each panel shows models of de(cid:12)nite values
of (cid:12) (rotation curve slope) and q ((cid:13)attening). The separation between models is
quite good. The triangles represent Rc = kpc, the squares represent Rc =
kpc, and the stars represent Rc = kpc. Panels marked E are for spherical
halos, while those marked E are for (cid:13)attened halos. Models with Rc = kpc
were also considered but they are easily distinguished since they typically have
(cid:28) (LMC)=(cid:28) (bul) > and therefore fall o(cid:11) the top of the (cid:12)gures.
Figure : Examples of LMC bt distributions for various model parameters. The
integral under each distribution is unity. The curve marked (a) is for a \standard"
spherical halo ((cid:12) = , q = , Rc = kpc, v = km/sec, R = : kpc, and
(cid:0) = : (cid:2) (cid:0) events/yr). Curve (b) has a shorter average duration ((cid:12) = (cid:0) :,
q = , Rc = kpc, v = km/sec, R = : kpc, and (cid:0) = : (cid:2) (cid:0)
events/yr). Curve (c) has a longer average duration ((cid:12) = :, q = :, Rc =
kpc, v = km/sec, R = : kpc, and (cid:0) = : (cid:2) (cid:0) events/yr). Finally
curve (d) has a maximal disk, which greatly reduces the amount of halo material
((cid:12) = , q = , Rc = kpc, v = km/sec, R = kpc, and (cid:0) = : (cid:2) (cid:0)
events/yr). The average of each distribution is (cid:10)bt(cid:11). All event rates scale (cid:0) / (cid:17)m.
|
astro-ph/0401367 | 1 | 0401 | 2004-01-19T16:27:38 | The origin of HI-deficiency in galaxies on the outskirts of the Virgo cluster. II. Companions and uncertainties in distances and deficiencies | [
"astro-ph"
] | The origin of the deficiency in neutral Hydrogen of 13 spiral galaxies lying in the outskirts of the Virgo cluster is reassessed. If these galaxies have passed through the core of the cluster, their interstellar gas should have been lost through ram pressure stripping by the hot X-ray emitting gas of the cluster. We analyze the positions of these HI-deficient and other spiral galaxies in velocity-distance plots, in which we include our compilation of velocity-distance data on 61 elliptical galaxies, and compare with simulated velocity-distance diagrams obtained from cosmological N-body simulations. We find that ~20% relative Tully-Fisher distance errors are consistent with the great majority of the spirals, except for a small number of objects, whose positions in the velocity-distance diagram suggest grossly incorrect distances, which implies that the Tully-Fisher error distribution function has non-gaussian wings. Moreover, we find that the distance errors may lead to an incorrect fitting of the Tolman-Bondi solution that can generate significant errors in the distance and especially the mass estimates of the cluster. We suggest 4 possibilities for the outlying HI-deficient spirals (in decreasing frequency): 1) they have large relative distance errors and are in fact close enough (at distances between 12.7 and 20.9 Mpc from us) to the cluster to have passed through its core and seen their gas removed by ram pressure stripping; 2) their gas is converted to stars by tidal interactions with other galaxies; 3) their gas is heated during recent mergers with smaller galaxies; and 4) they are in reality not HI-deficient (e.g. S0/a's misclassified as Sa's). | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. sanchis
(DOI: will be inserted by hand later)
October 29, 2018
The origin of H I-deficiency in galaxies on the outskirts of the
Virgo cluster
II. Companions and uncertainties in distances and deficiencies
T. Sanchis1, G. A. Mamon2
,
3, E. Salvador-Sol´e1
,
4, and J. M. Solanes1
,
4
1 Departament d'Astronomia i Meteorologia, Universitat de Barcelona, Mart´ı i Franqu`es 1, 08028 Barcelona,
Spain
e-mail: [email protected]; [email protected]; [email protected]
2 Institut d'Astrophysique de Paris (CNRS UMR 7095), 98 bis Bd Arago, F-75014 Paris, France
e-mail: [email protected]
3 GEPI (CNRS UMR 8111), Observatoire de Paris, F -- 92195 Meudon Cedex, France
4 CER on Astrophysics, Particle Physics, and Cosmology, Universitat de Barcelona, Mart´ı i Franqu`es 1, 08028
Barcelona, Spain
Received ............; accepted ...........
Abstract. The origin of the deficiency in neutral Hydrogen of 13 spiral galaxies lying in the outskirts of the Virgo
cluster is reassessed. If these galaxies have passed through the core of the cluster, their interstellar gas should have
been lost through ram pressure stripping by the hot X-ray emitting gas of the cluster. We analyze the positions
of these H I-deficient and other spiral galaxies in velocity-distance plots, in which we include our compilation of
velocity-distance data on 61 elliptical galaxies, and compare with simulated velocity-distance diagrams obtained
from cosmological N -body simulations. We find that ∼ 20% relative Tully-Fisher distance errors are consistent
with the great majority of the spirals, except for a small number of objects, whose positions in the velocity-distance
diagram suggest grossly incorrect distances, which implies that the Tully-Fisher error distribution function has
non-gaussian wings. Moreover, we find that the distance errors may lead to an incorrect fitting of the Tolman-
Bondi solution that can generate significant errors in the distance and especially the mass estimates of the cluster.
We suggest 4 possibilities for the outlying H I-deficient spirals (in decreasing frequency): 1) they have large relative
distance errors and are in fact close enough (at distances between 12.7 and 20.9 Mpc from us) to the cluster to
have passed through its core and seen their gas removed by ram pressure stripping; 2) their gas is converted to
stars by tidal interactions with other galaxies; 3) their gas is heated during recent mergers with smaller galaxies;
and 4) they are in reality not H I-deficient (e.g. S0/a's misclassified as Sa's).
Key words. Galaxies: clusters: individual: Virgo -- galaxies: evolution -- galaxies: ISM -- cosmology: distance scale
-- methods: N-body simulations
4
0
0
2
n
a
J
9
1
1
v
7
6
3
1
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1. Introduction
During the last thirty years, substantial work has been
carried out to quantify the effects of the environment on
the properties of galaxies, i.e. to distinguish the character-
istics of galaxies in dense, aggressive regions such as clus-
ters, especially cluster cores, and low density zones away
from clusters and groups. The nearby Virgo cluster is an
excellent candidate for such studies, because its proxim-
ity allows 1) an easy comparison with field galaxies and
2) the measurement of distances independent of radial ve-
locity, providing three-dimensional information on cluster
Send offprint requests to: T. Sanchis
properties (e.g. Young & Currie 1995; Gavazzi et al. 1999;
Solanes et al. 2002) impossible to obtain in other clusters.
One particular characteristic of spiral galaxies is their
neutral Hydrogen (H I) gas content. It has been known
for some time that a class of H I-deficient galaxies ex-
ists in the centers of clusters, such as the Virgo cluster
(Chamaraux, Balkowski, & G´erard 1980), and their defi-
ciency, normalized to their optical diameter and morpho-
logical type, is anti-correlated with their clustercentric dis-
tance (Haynes & Giovanelli 1986; Solanes et al. 2001).
Chamaraux et al.
(1980) considered ram pressure
stripping of
interstellar atomic Hydrogen by the hot
diffuse intracluster gas, first proposed by Gunn & Gott
2
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
(1972), as one cause of the H I deficiency pattern of Virgo
spirals. For galaxies moving face-on, ram pressure scales
as the intracluster gas density times the square of the
galaxy's velocity relative to this intracluster gas, so the
role of ram pressure stripping was strengthened by the
correlation of H I deficiency with the X-ray luminosity of
clusters found by Giovanelli & Haynes (1985; see however
Solanes et al. 2001) and by the presence of truncated H I
disks observed by Cayatte et al. (1990).
It hence came as a surprise when Solanes et al. (2002)
discovered spiral galaxies, whose Hydrogen content was
deficient relative to other spiral galaxies of the same mor-
phological type, far away from the core of Virgo, with
several deficient spirals in the foreground of Virgo and
particularly one group of spirals 7 to 10 Mpc (depend-
ing on the adopted distance to Virgo) behind the cluster
with one-third of its 15 spirals displaying H I deficiencies
deviating by more than 3 σ from normalcy.
In a followup study, Sanchis et al. (2002) explored the
possibility that the group of deficient spirals behind Virgo
might have plunged into the cluster core a few Gyr ago
and lost their gas by interacting with the hot intraclus-
ter medium. Sanchis et al. applied a simple dynamical
model that considered not only the classical deduction of
the virgocentric infalling regime, but also the trajectories
of galaxies that had already crossed once the center of
the core. This model succeeded in explaining the overall
characteristics of the observed velocity-distance diagram,
including the envelope of the virialized cluster drawn by
galaxies on first rebound. With the available data, they
argued, from a statistical point of view, that one could
not discard the possibility that some of the gas-deficient
objects, including the background group in the outskirts of
the Virgo Cluster Region, were not newcomers, but have
passed through the cluster and lost their gas by ram pres-
sure stripping.
We (Mamon et al. 2004) have recently estimated the
maximum radius out to where galaxies falling into a clus-
ter can bounce out. We found that this maximum rebound
radius corresponds to 1 to 2.5 cluster virial radii. After es-
timating the virial radius of the Virgo cluster using X-ray
observations, we concluded that the H I-deficient galaxies
located in the foreground or background of the Virgo clus-
ter appear to lie much further from the center of the clus-
ter than these maximum distances, and therefore that ram
pressure stripping cannot account for the H I-deficiency of
these outlying galaxies if their distances are correct.
In this paper, we investigate in further detail the ori-
gin of the H I-deficiency in the spiral galaxies that appear
too far from the Virgo cluster to have suffered ram pres-
sure stripping of their interstellar gas, by 1) adding to
our four-dimensional (angular position, distance and ra-
dial velocity) sample of spiral galaxies, our compilation of
the literature for elliptical galaxies towards Virgo, leading
us to identify more precisely the density of the environ-
ment of the background deficient galaxies and estimate
more accurately the distance to the center of Virgo, which
is a crucial ingredient for the dynamical infall model, and
2) using cosmological N -body simulations that have the
advantages of their lack of observational errors, complete-
ness, and of defining a realistic infall pattern.
The sample of elliptical galaxies is described in Sect. 2.
In Sect. 3, after describing the cosmological N -body sim-
ulations used here, we compare the simulated and obser-
vational samples to evaluate the influence of errors in dis-
tance estimates. Finally in Sect. 4, we discuss possible ex-
planations for the H I deficiency of the outlying galaxies.
2. The inclusion of ellipticals into the Virgo
galaxy sample
Using only late-type spirals, Sanchis et al. (2002) applied
a spherical infall model to the velocity field of the Virgo
Cluster Region and obtained two possible distances of
20 and 21 Mpc. Other distance estimates to the Virgo
cluster, based upon spiral galaxies, produced distances in
the range from 16 Mpc (Tully & Shaya 1984) to 21 Mpc
(Ekholm et al. 2000).
An alternative way to estimate the cluster distance
is by using the early-type components of the cluster, as
they are thought to be better tracers of the center of
the cluster due to morphological segregation (see e.g.
Ferguson & Binggeli 1994). To do so, we have collected
as many as possible early-type galaxies (E's, dE's and
S0's) with redshift-independent distance measurements.
Heliocentric velocities were extracted from the LEDA cata-
log of the HyperLEDA database and converted to the Local
Group reference frame, while the direct distant measure-
ments are obtained from a variety of sources.
Table 1. Datasets contributing to elliptical sample
Methoda Authorsb Band rel. accuracy Number
FP
L − n
R − n
SBF
SBF
SBF
SBF
GCLF
G
YL
YR
T
JI
JK
N
K
H
BJ
BJ
I
I
K ′
I
V −I
(rms)
0.41
0.54
0.54
0.18
0.07
0.17
0.15
0.07
41
9
14
31
6
7
15
8
a FP = Fundamental Plane; L − n, R − n = Luminosity /
S´ersic shape and Radius / S´ersic shape relations; SBF =
Surface Brightness Fluctuations; GCLF = Globular Cluster
Luminosity Function. b G = Gavazzi et al. (1999); T =
Tonry et al.
(2001); JI ,JK = Jensen, Tonry, & Luppino
(1998); N = Neilsen & Tsvetanov
(2000); K =
Kundu & Whitmore
(2001); YL,YR = Young & Currie
(1995).
Table 1 summarizes the distance datasets on the early-
type galaxies towards Virgo. Although different authors
use different methods with different accuracy, we have
not found systematic deviations between them, and fur-
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
3
thermore found that the spread in differences of the dis-
tance moduli was consistent with the quoted errors, for
all sets of early-type galaxies common to two authors.
Therefore, we estimated a global distance modulus d for
each galaxy, using the maximum likelihood estimator (e.g.
Bevington & Robinson 1992)
,
(1)
i
d = Pi di/σ2
Pi 1/σ2
i
where di are the distance moduli and σi are the abso-
lute errors on these distance moduli. Equation (1) sup-
poses that the relative error on distance measurements
is independent of distance and that the p.d.f. of the
distance moduli (proportional to log distance) is gaus-
sian. The error on the global distance modulus is (e.g.
Bevington & Robinson)
.
(2)
σ2
d
=
1
Pi 1/σ2
i
Although this estimator of the error on the distance mod-
ulus is independent on the scatter of the measurements,
we show in Appendix A (eqs. [A.7] and [A.8]) that equa-
tions (1) and (2) also represent the median and the half-
width (as defined by (x84 − x16)/2) of the likelihood func-
tion, respectively.
Table 2 provides our final sample of 63 early-type
galaxies. Columns (1), (2) and (3) give the name of the
galaxy, with its coordinates in columns (4) and (5). In col-
umn (6), we give the distance (in Mpc, from eq. [1]) and
its relative error (eq. [2]) in column (7). Column (8) gives
the velocity relative to the Local Group and column (9)
provides the sources for the distance.
Figure 1 shows a wedge diagram of the Virgo region
with the 61 early-types shown as circles and the 146 spirals
shown as triangles. One clearly notices the concentration
of elliptical galaxies at a distance of D ≃ 17 Mpc, which
is spread out in velocity space.
3. Observed and simulated velocity-distance
relation towards the Virgo Cluster
For a better understanding of the kinematics of galaxies
in the direction of the Virgo cluster, we have simulated an
observation of the velocity-distance relation using the dark
matter particles in the GalICS cosmological simulation.
3.1. N -body simulations
The N -body simulations used here to find N -body replica
of the Virgo cluster were carried out by Ninin (1999), as
described by Hatton et al. (2003) in the context of the
GalICS hybrid N -body/semi-analytic model of hierarchi-
cal galaxy formation. Here, we are only interested in the
density and velocity fields directly traced by dark matter
particles. The N -body simulation contains 2563 particles
of mass 8.3 × 109M⊙ in a box of 150 Mpc size and it is
run with a softening length amounting to a spatial resolu-
tion of 29 kpc. The simulation was run for a flat universe
with cosmological parameters Ω0 = 0.333, ΩΛ = 0.667,
H0 = 66.7 km s−1 Mpc−1, and σ8 = 0.88. Once the sim-
ulation is run, halos of dark matter are detected with a
'Friends-of-Friends' (FoF) algorithm (Davis et al. 1985),
with a variable linking length such that the minimum mass
of the FoF groups is 1.65 × 1011M⊙ (20 particles) at any
time step. With this method, over 2 × 104 halos are de-
tected at the final timestep, corresponding to the present-
day (z = 0) Universe. The GalICS halo finder does not
allow halos within halos, so that a cluster, to which is as-
signed a massive halo, cannot contain smaller halos within
it.
3.2. Picking isolated halos
Among the 12 halos that GalICS produces at z = 0 with
MFoF > 1014 M⊙, we wish to choose those that resemble
the Virgo cluster, both in terms of mass and environment.
As we can see in the radial phase space diagrams
of dark matter halos of the simulations (Fig. 1 of
Mamon et al. 2004),
it is quite common to find small
groups around the main cluster halo, whereas the Virgo
cluster is not known to have massive neighbors, except for
the M49 group just within the virial radius. Note that the
size of the simulation box, L = 150 Mpc, implies, given
the periodic boundary condition, a maximum separation
between halos of √3 (L/2) = 130 Mpc.
The effect of a neighboring halo will be from its tide
on the test halo. The ratio of the tidal acceleration on a
particle of the test halo at distance r from its center caused
by a neighboring halo at distance R from the center of
the first halo to the acceleration of this particle by the
potential of its halo amounts to a mean density criterion:
G M2(R)
R3
r
atid ≈
and
G M1(r)
r2
ahalo =
so that
atid
ahalo ≈
ρ2(R)
ρ1(r)
,
(3)
(4)
(5)
where the densities are average (and not local) and sub-
scripts 1 and 2 stand for the central and the neighboring
halo respectively. For clarity, we denote r1, M1, r2 and
M2 the virial radii and masses for the test and neighbor-
ing halo, respectively (i.e. we drop the '100' subscript).
Then, equating r = r1, noting that the mean density at
the virial radius is the same for each halo, i.e. 100 times
the critical density of the Universe, and writing
η =
R
r2
,
(6)
the influence of each neighbor is measured by the mean
density of the neighbor measured at the center of the test
4
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Table 2. Final sample of early-type galaxies
VCC
49
220
312
319
342
345
355
369
575
648
685
731
763
778
784
810
828
881
940
953
965
1025
1030
1146
1173
1196
1226
1231
1242
1250
1253
1279
1297
1308
1316
1321
1386
1412
1420
1489
1535
1537
1539
1549
1619
1630
1632
1664
1669
1692
1720
1834
1869
1883
1902
1903
Galaxy
NGC
4168
4233
4255
--
4259
4261
4262
4267
4318
4339
4350
4365
4374
4377
4379
--
4387
4406
--
--
--
4434
4435
4458
--
4468
4472
4473
4474
4476
4477
4478
4486B
--
4486
4489
--
4503
--
--
4526
4528
--
--
4550
4551
4552
4564
--
4570
4578
4600
4608
4612
4620
4621
Messier
RA
Dec
(J2000)
--
--
--
--
--
--
--
--
--
--
--
--
84
--
--
--
--
86
--
--
--
--
--
--
--
--
49
--
--
--
--
--
--
--
87
--
--
--
--
--
--
--
--
--
--
--
89
--
--
--
--
--
--
--
--
59
12h12m17.s3
12h17m07.s7
12h18m56.s1
12h19m02.s0
12h19m22.s2
12h19m23.s2
12h19m30.s6
12h19m45.s4
12h22m43.s3
12h23m34.s9
12h23m57.s7
12h24m28.s3
12h25m03.s7
12h25m12.s3
12h25m14.s8
12h25m33.s8
12h25m41.s8
12h26m12.s2
12h26m47.s1
12h26m54.s7
12h27m03.s1
12h27m36.s7
12h27m40.s6
12h28m57.s6
12h29m14.s8
12h29m30.s9
12h29m46.s7
12h29m48.s9
12h29m53.s7
12h29m59.s1
12h30m02.s2
12h30m17.s5
12h30m32.s0
12h30m45.s9
12h30m49.s4
12h30m52.s3
12h31m51.s5
12h32m06.s3
12h32m12.s3
12h33m14.s0
12h34m03.s1
12h34m06.s2
12h34m06.s3
12h34m14.s9
12h35m30.s7
12h35m38.s1
12h35m40.s0
12h36m27.s0
12h36m30.s6
12h36m53.s5
12h37m30.s6
12h40m23.s0
12h41m13.s4
12h41m32.s8
12h41m59.s4
12h42m02.s3
+13◦12′17′′
+07◦37′26′′
+04◦47′09′′
+13◦58′48′′
+05◦22′34′′
+05◦49′32′′
+14◦52′38′′
+12◦47′53′′
+08◦11′52′′
+06◦04′54′′
+16◦41′33′′
+07◦19′04′′
+12◦53′13′′
+14◦45′43′′
+15◦36′26′′
+13◦13′30′′
+12◦48′35′′
+12◦56′44′′
+12◦27′14′′
+13◦33′57′′
+12◦33′39′′
+08◦09′14′′
+13◦04′44′′
+13◦14′29′′
+12◦58′41′′
+14◦02′55′′
+07◦59′59′′
+13◦25′46′′
+14◦04′06′′
+12◦20′55′′
+13◦38′11′′
+12◦19′40′′
+12◦29′27′′
+11◦20′36′′
+12◦23′28′′
+16◦45′30′′
+12◦39′24′′
+11◦10′35′′
+12◦03′42′′
+10◦55′43′′
+07◦41′57′′
+11◦19′15′′
+12◦44′40′′
+11◦04′16′′
+12◦13′13′′
+12◦15′49′′
+12◦33′22′′
+11◦26′18′′
+13◦38′18′′
+07◦14′46′′
+09◦33′16′′
+03◦07′02′′
+10◦09′18′′
+07◦18′50′′
+12◦56′32′′
+11◦38′45′′
D
(Mpc)
34.20
32.96
28.71
7.59
57.02
32.27
14.86
14.66
25.35
16.44
14.86
22.07
17.71
15.49
14.19
12.36
17.21
17.40
7.24
18.88
9.29
25.50
13.12
17.97
14.96
15.85
16.27
17.12
15.14
19.08
16.44
16.68
16.31
17.86
16.82
17.04
7.08
11.75
11.75
11.30
16.61
14.52
8.83
19.32
17.84
17.24
15.31
14.86
8.79
13.80
17.94
7.35
20.23
14.06
21.28
14.80
rel. error
v
sources
(km s−1)
0.41
0.41
0.41
0.38
0.41
0.16
0.41
0.41
0.41
0.17
0.41
0.05
0.07
0.41
0.38
0.38
0.35
0.04
0.54
0.38
0.54
0.15
0.41
0.09
0.38
0.13
0.03
0.05
0.41
0.11
0.41
0.09
0.03
0.38
0.09
0.13
0.38
0.41
0.38
0.38
0.17
0.41
0.54
0.54
0.09
0.15
0.04
0.15
0.54
0.41
0.06
0.20
0.41
0.41
0.28
0.05
2128
2181
1603
−266
2306
2012
1181
852
1047
1108
1049
1058
733
1182
871
−516
373
−465
1337
−678
614
897
611
499
2292
732
693
2041
1405
1767
1164
1220
136
1511
1095
772
1135
1184
847
−94
427
1195
1216
1202
208
973
115
947
427
1559
2102
615
1633
1697
1012
258
G
G
G
YLYR
G
GT
G
G
G
T
G
JI JK NT
GNT
G
T
YLYR
GT
JI JK KNT
YR
YLYR
YR
GT
G
GKNT
YLYR
T
GJI JK KNT
GKNT
G
NT
G
GNT
KN
YLYR
GNT
JK T
YLYR
G
YLYR
YLYR
GT
G
YR
YR
KNT
GT
GJI JK KNT
GT
YR
G
GJI JK T
T
G
G
T
GKNT
Table 2. Final sample of early-type galaxies (ctd)
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
5
VCC
1938
1939
1978
2000
2087
2092
2095
Galaxy
NGC
4638
4624
4649
4660
4733
4754
4762
Messier
RA
Dec
(J2000)
--
--
60
--
--
--
--
12h42m47.s5
12h42m50.s0
12h43m40.s1
12h44m32.s0
12h51m06.s8
12h52m17.s5
12h52m55.s9
+11◦26′32′′
+02◦41′16′′
+11◦33′08′′
+11◦11′24′′
+10◦54′43′′
+11◦18′50′′
+11◦13′49′′
D
(Mpc)
17.33
15.14
16.36
15.66
14.93
16.94
10.81
rel. error
v
sources
(km s−1)
0.21
0.06
0.04
0.11
0.18
0.12
0.41
956
924
969
914
772
1206
818
GT
GJI JK T
GKNT
GNT
T
GT
G
Fig. 1. Positions of 146 spiral (triangles) and 61 elliptical (circles) galaxies in the Virgo field (within 9◦ from M87)
as a function of Declination and distance (assumed correct, top) and radial velocity (bottom). H I-deficient galaxies
(deviating more than 3σ from normalcy) are shown as filled triangles -- with the ones apparently outside the cluster
(Table 7, below) shown as surrounded filled triangles (red triangles in the Electronic version of the Journal), while
Messier ellipticals (see Table 2) are shown as filled circles, with M49 to the South, and M59, M89, M60, M87, M86
and M84 at increasing distances (top plot) and M89, M59, M84, M60 and M87 at increasing radial velocities (bottom
plot, M86 has v < 0).
halo
ρ2(R) =
∝
3 M2(R)
4 π R3
M2(R)/M2
η3
= (cid:8)η3 [ln η − η/(1 + η)](cid:9)−1
,
(7)
where we assumed NFW density profiles and where the
last equality is from Cole & Lacey (1996). Since the mean
density is a monotonously decreasing function of radius
(i.e. the function of η in eq. [7] decreases with η), the
most isolated halos will have the highest minimum values
of η.
The output from GalICS includes the mass of each halo
obtained from the FoF estimator, so that one can easily
perform a first guess of which of the halos with mass close
to that of Virgo are as isolated as appears the Virgo clus-
ter. Table 3 lists for the 10 most massive halos in the
6
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
simulation, the parameters of its most perturbing neigh-
bor (that with the lowest value of η) using FoF quantities
for radii and masses. Columns 2 to 4 list the virial radius,
mass and circular velocity obtained using the more precise
spherical overdensity (SO) method, which allow a direct
comparison with the value we obtained for the Virgo clus-
ter in Mamon et al. (2004).
Table 3. Most perturbing neighbors of 10 most massive
halos in GalICS simulation
add the Hubble flow. Therefore:
D = Ds
v = us
rV
rs
vV
vs
+ H0 D
(8)
(9)
where us represents the comoving radial velocity of the
particle in the simulation with respect to the observer in
comoving units.
Inverting equation (8), we need to place the observer
at a distance
ηmin
Dobs = DV
rs
rV
(10)
rSO
1
rank
(FoF) (Mpc)
M SO
1
vSO
1
(1014 M⊙)
(km s−1)
1
2
3
4
5
6
7
8
9
10
3.19
2.42
2.50
2.54
2.10
2.00
1.97
1.78
1.84
2.03
16.7
7.4
8.1
8.5
4.8
4.1
3.9
2.9
3.2
4.3
1503
1143
1180
1200
989
942
927
842
869
958
1
2
R M FoF
1 M FoF
rSO
0.011
0.55
0.003
0.58
0.55
0.020
0.011
0.53
0.001
0.62
0.001
0.58
0.57
0.031
0.002
0.61
0.103
0.58
0.51
0.003
3.7
3.1
2.4
2.4
2.9
2.5
3.4
2.5
2.2
2.6
to the halo used to mimic the Virgo cluster (in simula-
tion units). Once we have Dobs, we can select different
observers by choosing different positions on a sphere of
radius Dobs centered on the halo and assigning to each
position, the mean velocity of the 10 nearest halos to the
observer, so as to give the observer an appropriate peculiar
velocity. As our galaxies in the Virgo cluster extend up to
Dmax = 50 Mpc, we will need to use particles around the
chosen dark matter halo up to a distance from the cluster
of Dmax − DV ≃ 20 r100.
It is interesting to check the peculiar velocities (or
equivalently the systemic velocities) found for the simu-
lated observer relative to the simulated cluster and com-
pare to that of our Local Group of galaxies relative to
the Virgo cluster, after rescaling the simulation radial and
velocity separations to the Virgo cluster attributes as in
equations (8) and (9). For 500 observers placed at a given
distance from the halo, but in different random directions,
we find a systemic velocity between the halo and the ob-
server (i.e. the mean of 10 nearest halos) in Virgo units
of 998 ± 118 km s−1 (errors are the 1 σ dispersion) when
we assume DV = 16.8 Mpc (as suggested by the posi-
tions of the early-type galaxies in Fig. 1, and especially
of M87, which lies at the center of the X-ray emission)
and 1308 ± 155 km s−1 when we assume DV = 21 Mpc,
as in Sanchis et al. (2002). The values of the systemic ve-
locity of the Virgo cluster in the literature range approx-
imately between 900 and 1000 km s−1 (Teerikorpi et al.
1992). Therefore, putting the Virgo cluster at 16.8 Mpc, as
indicated by the early type galaxies, gives a better match
to the systemic velocity of the cluster. In addition, if we
calculate the number of all galaxies (early-type and late-
type ones) per distance bin, we found the maximum of
the distribution between 16 − 18 Mpc. Even if our sample
lacks of completeness, this result, together with the pre-
vious ones, lead us to choose DV = 16.8 Mpc for the com-
parison between the N -body simulations and the Virgo
cluster.
3.4. Velocity-distance relation without distance errors
We take all the objects lying within a cone of half-opening
9◦ aligned with the axis connecting the observer to the
center of the halo.
As seen in Table 3, the most perturbing halo is always
inside r1. Therefore, it is not clear that we can easily es-
timate the exact virial radius and mass of each perturb-
ing halo with the same spherical overdensity method as
in Mamon et al. (2004) to produce a better estimate of
isolated halos. As we expect rSO
to be approximately pro-
portional to rFoF
, we take the results of Table 3 as good
indicators of the degree of isolation of halos in the simu-
lations.
1
1
As shown in Table 3, the most isolated halo in the
simulations is halo 1, the most massive halo. Picking the
most massive halo also ensures that no other massive halos
will distort the mock velocity-distance relation.
3.3. Rescaling the simulations to the scales of the
Virgo cluster
We now need to rescale the simulated clusters to Virgo.
A reasonable way to put two halos in comparable units is
to normalize distances with the virial radius, r100, and
velocities with the circular velocity of the halo at the
virial radius, v100. We therefore convert from the sim-
ulation frame to the Virgo cluster frame by rescaling
the terms related to distance with the factor (omitting
the '100' subscripts for clarity) rV/rs and the terms re-
lated to velocities by the analogous factor vV/vs, where
superscripts 's' and 'V' refer to the simulation and the
Virgo cluster, respectively. We adopt the Virgo cluster
virial radius and circular velocity derived by Mamon et al.
(2004) from X-ray observations rV = 1.65 h−1
2/3 Mpc, where
h2/3 = H0/(66.7 km s−1 Mpc−1), and vV = 780 km s−1.
Moreover, the N -body simulations give us comoving
velocities us of dark matter particles, to which we must
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
7
Because observed galaxy catalogs, such as the VCC
(Binggeli, Sandage, & Tammann 1985) catalog of Virgo
cluster galaxies, are magnitude-limited, we imposed upon
our simulated catalog (of dark matter particles) the
same magnitude limit as the VCC, namely B < 18
(Binggeli et al. 1985). For this, we consider bins of dis-
tance to the observer, and, in each distance bin, we re-
move a fraction of dark matter particles equal to the pre-
dicted fraction of particles fainter than B = 18 given the
distance and the gaussian-shaped luminosity function of
spiral galaxies derived by Sandage, Binggeli, & Tammann
(1985). This method supposes that the luminosity func-
tion is independent of environment (i.e. distance to the
center of the Virgo cluster) and that those VCC galax-
ies with redshift-independent distance measurements have
the same magnitude limit as the VCC galaxies in general.
This luminosity incompleteness is only noticeable at large
distances, where it amounts to reducing the number of
particles by less than 10%.
Figure 2 shows the velocity-distance plot of the sim-
ulated dark matter particles (small points), rescaled to
Virgo units, and from which particles statistically fainter
than the magnitude limit of the observed catalogs were re-
moved. Superposed are the spiral galaxies (triangles) from
Solanes et al. (2002) and the early-type galaxies (circles)
compiled in Table 2. We have added to our sample the
spiral galaxies compiled by Solanes et al. (2002) with reli-
able Tully-Fisher (Tully & Fisher 1977, hereafter TF) dis-
tances but no H I data. We list these galaxies in Table 4,
with their coordinates, morphological type (from LEDA),
systemic velocity and the final distance with errors as-
signed to it by Solanes et al. (2002), who estimated errors
in distance as the mean deviation between the rescaled
mean distance and the different estimates used (with this
method, galaxies with only one estimator cannot be as-
signed an error and their distances are considered uncer-
tain). We represent these objects as squares in Figure 2.
Although Figure 2 concerns a single observer and a sin-
gle halo to represent the Virgo cluster (i.e. halo 1, which is
the most isolated), the results are very similar for all halos
studied as well as for any observer whose velocity with re-
spect to the center of the main halo is similar to the mean
of the 500 random trials. In other words, Figure 2 provides
a realistic general velocity-distance relation for a Virgo-
like cluster in a flat Universe with a cosmological con-
stant. Very recently, Klypin et al. (2003) have produced a
similar plot for dark matter particles from their adaptive
mesh refinement simulation. Note that the groups visible
in the 3D phase space plot (Fig. 1 in Mamon et al. 2004)
are much less visible here and the projection places them
at smaller distances from the center of the main cluster.
Figure 2 indicates that the velocity field drawn by
the observed spirals is noticeably different from the one
followed by the dark matter particles of the simulation.
Nevertheless, the figure shows that 1) the center of the
halo corresponds to the location of greater galaxy density
(not only for the early-type galaxies used to mark the cen-
ter of the Virgo cluster), 2) the thickness of the halo in
velocity space (the velocity dispersion) corresponds to the
corresponding thickness of the central observed galaxies,
and 3) the majority of the elliptical galaxies outside the
cluster also appear to fit well the location of the particles
in the outer infalling/expanding locus. These 3 facts sug-
gest that both the distance to the halo and the rescaling
factors are well chosen and that the differences between
the Virgo cluster and the halo in the simulation should be
attributable to errors arising from the distance estimates
(as the velocity errors are negligible). It appears unlikely
that the galaxies that, for a given distance, lie within more
than 1000 km s−1 from the locus of dark matter particles
belong to the low- or high-velocity tails of large velocity
dispersion groups near the cluster, since in our simula-
tions, there are no such high velocity dispersion groups
outside the main cluster.
3.5. Velocity-distance relation with distance errors
Given that most of the galaxies outside the velocity field
drawn by dark matter particles are spirals, we now build
a new velocity-distance diagram of our simulated particles
incorporating the mean relative distance error for the spi-
ral galaxy sample. Solanes et al. (2002) combined different
studies of the TF relation to calculate the distances of spi-
rals in the Virgo cluster. These studies give dispersions of
the TF relation of roughly 0.4 mag., corresponding to an
uncertainty of 18% in relative distance. Figure 3 shows
the corresponding velocity-distance plot for our simulated
Virgo line of sight, with inclusion of gaussian relative dis-
tance errors with σ ln D = 0.20 (a slightly more conserva-
tive value than 0.18). We omit particles beyond 40 Mpc,
which corresponds to the distance (before folding in the
distance errors) where the particles have a systemic veloc-
ity approximately equal to the maximum of the observed
galaxy sample.
Figure 3 indicates that the inclusion of distance errors,
at the 20% relative rms level, allows us to reproduce fairly
well the observed velocity-distance diagram. There are still
several galaxies that lie far outside the general locus of the
dark matter particles. Among them are two galaxies (VCC
319, which is part of a galaxy pair and M90 = NGC 4569
= Arp 76, which is tidally perturbed by a close neighbor)
apparently in the foreground of the cluster between 7 and
10 Mpc and with v < −200 km s−1, and four galaxies (IC
3094, NGC 4299, which is part of a pair, NGC 4253 and
VCC 1644) in the background between 34 and 42 Mpc
and v < 800 km s−1. Inspection of Figure 2 indicates that
these four galaxies are at more than 2000 km s−1 from the
expected locus of the Tolman-Bondi pattern given negligi-
ble distance errors. It is highly unlikely that the peculiar
velocities in interacting pairs can reach values as high as
2000 km s−1, hence the distances to these galaxies appear
to be grossly incorrect.
We list the most discrepant galaxies in Table 5. Most
of these galaxies have a single distance measurement (the
elliptical VCC 319 has a single S´ersic fit distance esti-
8
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Table 4. Spiral galaxies without H I data
Galaxy
VCC NGC
132
343
912
952
1605
1673
1933
--
--
4413
--
--
4567
--
RA
Dec
(J2000)
12h15m03.s9 +13◦01′55′′
12h19m22.s0 +07◦52′16′′
12h26m32.s2 +12◦36′39′′
12h26m55.s5 +09◦52′58′′
12h35m14.s5 +10◦25′53′′
12h36m32.s9 +11◦15′28′′
12h42m44.s7 +07◦20′16′′
T
8
8
2
5
5
4
2
v
(km s−1)
1965
2336
−13
853
951
2133
2271
D
(Mpc)
10.76+2.60
−2.10
23.77
15.63+2.57
−2.20
4.77+4.60
−3.88
24.66+6.82
−5.34
26.30
35.16+0.33
−0.32
Fig. 2. Simulated and observed velocity-distance diagrams. The observer is at (0, 0) and sees a cone of angular radius
9◦. Dots represent the velocity field traced by the particles (1 in 50 for clarity) in the cosmological N -body simulation
in the direction of the most massive halo, for which distances and velocities have been rescaled (eqs. [8] and [9])
incorporating a Hubble flow with H0 = 66.7 km s−1 Mpc−1, as in the simulation. Superposed are galaxies of the Virgo
cluster. Circles and triangles represent early-type and late-type galaxies, respectively. The size of the triangles is
proportional to the H I-deficiency of the spiral galaxies measured in units of the mean standard deviation for field
objects (= 0.24). Squares represent additional spiral galaxies with no H I-deficiency data. Open symbols indicate
galaxies with uncertain distances (with distance errors greater than 5 Mpc or spirals with only one distance estimate).
Dashed lines show the unperturbed Hubble flow with H0 = 66.7 and 70 km s−1 Mpc−1, respectively (going upwards).
The locations of the vertical spikes of simulated particles away from the Virgo cluster (assumed to lie at 16.8 Mpc)
are not meaningful and should not be compared to the measured galaxy distances.
mator, which is known to be inaccurate -- see Table 1).
However, it is a surprise to find a bright object such as
NGC 4569 (M90) in this list, especially since it has 6 differ-
ent Tully-Fisher distance measurements, 5 of which yield a
distance between 7.9 and 10.2 Mpc and one (Gavazzi et al.
1999) giving D = 16.1 Mpc, which is what we expect from
Figure 2.
Since the 20% relative errors correctly reproduce the
envelope of particles from the cosmological simulation, in-
creasing the relative rms error to say 30% will not explain
the distances to these discrepant spirals. We therefore con-
clude that the distances of the galaxies listed in Table 5
must be highly inaccurate.
In their dynamical, Tolman-Bondi, modeling of the
Virgo region, Sanchis et al. (2002) only used those spiral
galaxies believed to trace the outer envelope of galaxies
expanding out of the Virgo cluster. Figure 2 clearly in-
dicates that the galaxies used by Sanchis et al. (roughly
aligned between distances of 34 and 45 Mpc) lie at
typically 600 km s−1 lower velocities for their distance
than predicted by our simulated velocity-distance model.
Figure 3 suggests that the alignment of these galaxies in
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
9
Fig. 3. Same as Figure 2, but incorporating gaussian relative distance errors of σ ln D = 0.2 for the dark matter
particles.
Table 5. Grossly incorrect galaxy distances
Galaxy
VCC NGC
RA
Dec
(J2000)
213 -- 12h16m56.s0 +13◦37′33′′
319 -- 12h19m01.s7 +13◦58′56′′
12h21m40.s9 +11◦30′03′′
491 4299
12h33m48.s0 +15◦10′05′′
1524 4523
12h35m51.s8 +13◦51′33′′
1644
12h36m49.s8 +13◦09′48′′
1690 4569
T
5
−4
8
9
9
2
v
D
(km s−1) Mpc
40.9
−278
7.6
−256
34.8
112
158
36.1
41.5
646
−328
9.4
VCC 1690 = NGC 4569 = M90 = Arp 76.
the velocity-distance plot may be fortuitous, given the
∼ 20% relative distance errors. This fortuitous alignment
forced the model of Sanchis et al. to shift the center of the
Virgo cluster to greater distances.
3.6. Tolman-Bondi estimates of the Virgo cluster mass
and distance
tograms of the number of pseudo-galaxies per 2 Mpc dis-
tance bin. Such histograms peak at the core of the halo,
in a region of width ∼ 4 Mpc for the no error set and up
to ∼ 12 Mpc for the 30% error set. We avoid the virialized
core by only considering pseudo-galaxies lying on the back
side of the identified cluster core.
In the Tolman-Bondi model used by Sanchis et al.,
galaxies are treated as test particles moving in a point-
mass gravitational potential that is independent of time.
The model velocity-distance curve is a function of the age
of the Universe, t0, the distance to the cluster, DV, the
point mass of the cluster, MV, and the systemic velocity
of the cluster relative to the Local Group of galaxies, vV.
Only three of these four parameters are independent, but
given the poor current velocity-distance data, it is bet-
ter to use only two free parameters and fix the third one.
Following Sanchis et al., we have fixed vV to the mean ve-
locity of the main halo for the 50 realizations (998 km s−1
for a distance of 16.8 Mpc, see section 3) and we have
allowed MV and DV to vary freely. The resulting t0 will
help us to see the viability of the results.
We now use the cosmological N -body simulations to quan-
tify the errors on the distance and mass to the Virgo
cluster obtained by fitting a Tolman-Bondi model to the
velocity-distance data outside the cluster core. We con-
struct simulated pseudo-galaxy samples by randomly se-
lecting as many points from the cosmological N -body sim-
ulation as there are galaxies in the Virgo cluster sample.
We apply fixed relative distance errors in the range 0% --
30% and, for each relative distance error, we construct
50 realizations of the Virgo sample. We identify pseudo-
galaxies presumably belonging to the core by building his-
Table 6 shows the results of the fits to the point-mass
Tolman-Bondi relation (see Sanchis et al. 2002) for the
different sets. For zero distance errors, the model repro-
duces fairly well the input data of DV = 16.8 Mpc and
MV = 1014.35 M⊙ (Mamon et al. 2004), as well as the age
of the Universe: t0 = 13.7 Gyr from the GalICS cosmologi-
cal simulations (which happens to perfectly match the age
recently obtained by Spergel et al. 2003 from the WMAP
CMB experiment). The Table also clearly indicates that
the errors in distance contribute to a slightly larger dis-
tance to the cluster (by up to 2%), combined with a much
10
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Table 6. Best fitting parameters for each dataset
Error
DV
(Mpc)
log(MV)
(M⊙)
t0
(Gyr)
mean
16.7
16.8
16.9
17.0
17.1
17.1
17.1
σ
0.5
0.5
0.6
0.8
0.9
1.0
1.0
0%
5%
10%
15%
20%
25%
30%
mean
14.5
14.6
14.9
15.1
15.5
15.7
16.1
σ
0.2
0.3
0.4
0.5
0.5
0.4
0.5
14.1
13.9
12.8
12.1
10.1
9.0
6.7
larger mass for the cluster (up to a factor 40!). This dra-
matic rise in the mass decreases the resulting t0 to unac-
ceptably low values. Therefore, distance errors can cause
the fitting of Tolman-Bondi solutions to the velocity vs.
distance of galaxies to yield incorrect cluster parameters.
For the case of real data on the Virgo cluster, the
same fitting procedure leads to MV ∼ 3 × 1016 M⊙ and
DV ∼ 17.4 Mpc, which imply t0 = 5.2 Gyr. Note that
the present fitting procedure takes into account all galax-
ies beyond the identified core of the cluster, whereas
Sanchis et al. only used those galaxies that are aligned be-
tween distances of 34 and 45 Mpc in the velocity-distance
plot, deducing a very different distance to the Virgo clus-
ter (between 20 and 21 Mpc). The comparison of the val-
ues of t0 from the real data and from the simulated data
convolved with different relative distance errors, suggests
that the TF distances may have up to 30% relative errors.
As shown above, the estimation of the distance to the
cluster provided by the point mass model is relatively good
even for appreciable errors in relative distances. Therefore,
one could still obtain a reasonable value for the mass of the
cluster by taking DV from the fit (i.e., 17.4 Mpc), together
with vV = 980 km s−1 (Teerikorpi et al. 1992) and t0 =
13.7 Gyr (Spergel et al. 2003) as input parameters for the
model and computing MV as a result. In this way, we
obtain a value of MV = 7 × 1014 M⊙. Notice that the
effective mass of the point mass model does not have to
coincide with the virial mass of the cluster, as the latter
one represents only the virialized part of the cluster, while
the former gives us an idea of the dynamical mass of the
whole Virgo region.
4. Discussion
As mentioned in Sec. 1 and clearly seen in Figure 2, there
is a non negligible number of H I-deficient galaxies distant
from the cluster core, in particular at distances of 10 and
28 Mpc from the Local Group. We now examine different
explanations for the presence of these H I-deficient galaxies
outside the core of Virgo:
1. very inaccurate distances (perhaps from the inade-
quacy of the Tully Fisher relation for galaxies under-
going ram pressure or tidal stripping) for galaxies that
are in fact within the Virgo cluster (in the core or
more probably in a rebound orbit) and which would
have lost some of their gas by ram pressure stripping
from the intracluster gas;
2. tidal interactions from members of a same group;
3. tidal interactions from neighbors;
4. recent minor mergers heating the gas;
5. overestimated H I-deficiency in early-type spirals.
In Table
compilation
isfy
at
7
of
least
listed all
are
Solanes et al.
one
the
of
galaxies
(2002) which
following
from the
sat-
criteria:
θ
the
to M87
angular distance
• D < 13 Mpc ,
• 21 < D < 50 Mpc ,
• θ/θ100 > 1 ,
where
and
is
θ100 = r100/DV = 5.◦6 is the angular virial radius
of the Virgo cluster. Columns (1) and (2) give the galaxy
names, the coordinates in columns (3) and (4), the galaxy
type, total blue extinction corrected magnitude and
optical diameter (all three from LEDA) are respectively
in columns (5), (6), and (7), the projected distance to
M87 in units of the virial radius (r100) of the cluster in
column (8), the velocity relative to the Local Group taken
from the Arecibo General Catalog (a private dataset
maintained by R. Giovanelli and M. P. Haynes at the
University of Cornell), in column (9) and column (10)
gives the adopted distance (Solanes et al. 2002). The last
3 columns of Table 7 are explained in Sect. 4.1.
4.1. Are the outlying H I-deficient galaxies truly too far
from the Virgo cluster to have had their gas ram
pressure stripped?
If one wishes to explain the outlying H I-deficient galaxies
as galaxies that have lost their interstellar gas by its strip-
ping by ram pressure from the hot intracluster gas, then
these galaxies must have passed through the core of the
Virgo cluster where the intracluster gas is dense and the
galaxy velocities large so that the ram pressure is largest.
Mamon et al. (2004) found that objects that have
passed through the core of a structure in the past can-
not be at distances greater than 1 -- 2.5 virial radii from
the cluster today. With their estimate of the distance (16.8
Mpc) and the virial radius (1.65 Mpc) of Virgo, this means
that galaxies that have passed through the core of Virgo
cannot lie at a distance from the Local Group greater than
18.5 or 20.9 Mpc (for rreb/r100 = 1 or 2.5, respectively)
nor can they lie closer than 15.1 or 12.7 Mpc from the
Local Group.
For the 3 H I-deficient spirals at 28 Mpc to have passed
through the core of Virgo, one would require that their dis-
tances be each overestimated by 52% (34%), which corre-
sponds to 2.3 σ (1.6 σ) events for their distance moduli, for
rreb/r100 = 1 (2.5), respectively. Such errors are possible,
as shown in Sect. 3.5. Note that none of the distances of
Table 7. H I-deficient galaxies possibly outside the Virgo cluster
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
11
Galaxy
RA
Dec
T
B c
T
D25
θ
v
D
N
P1
P2.5
VCC
--
522
524
559
713
979
1043
1330
1569
1690
1730
1760
1859
NGC
4064
4305
4307
4312
4356
4424
4438
4492
--
4569a
4580
4586
4606
(J2000)
12h04m11.s8
12h22m03.s5
12h22m06.s3
12h22m32.s0
12h24m15.s9
12h27m13.s3
12h27m46.s3
12h30m58.s9
12h34m31.s4
12h36m50.s5
12h37m49.s5
12h38m28.s1
12h40m57.s7
18◦26′33′′
12◦44′27′′
09◦02′27′′
15◦32′20′′
08◦32′10′′
09◦25′13′′
13◦00′30′′
08◦04′41′′
13◦30′23′′
13◦09′54′′
05◦22′09′′
04◦19′09′′
11◦54′46′′
1
1
3
2
6
1
1
1
5
2
2
1
1
11.73
12.87
11.84
11.76
13.02
11.99
10.55
13.04
14.62
9.63
12.39
12.06
12.28
4.′0
2.′0
3.′5
4.′7
2.′6
3.′4
8.′7
1.′9
0.′8
10.′4
2.′0
3.′9
2.′9
(km s−1) (Mpc)
9.8
40.7
27.4
10.8
29.4
4.1
10.4
28.1
23.3
9.4
19.3
14.4
12.7
837
1814
913
47
998
314
-- 45
1638
687
-- 328
893
639
1528
8.◦8
2.◦2
4.◦0
3.◦7
4.◦2
3.◦1
1.◦0
4.◦3
1.◦4
1.◦7
7.◦2
8.◦3
2.◦5
942
200
1846
754
1377
106
755
1261
8967
143
8724
4657
7648
0.00
0.00
0.41
0.95
0.19
0.00
1.00
0.08
0.72
1.00
0.00
0.00
0.95
0.81
0.00
0.48
1.00
0.23
0.00
1.00
0.11
0.94
1.00
0.84
0.98
1.00
a NGC 4569 is also M90.
the 13 galaxies listed in Table 7 are based upon Cepheid
measurements, which are much more precise than Tully-
Fisher estimates. On the other hand, given the presence in
our sample of some spirals with grossly incorrect distances
(see Sect. 3.5), one may wonder if the distances to the 3
spirals at 28 Mpc could also be grossly incorrect: they
could lie at the cluster distance of 16.8 Mpc, thus leading
in this case to distances overestimated by 66%, which for
20% rms relative errors correspond to three 2.8 σ events for
their distance moduli, which appears unlikely, unless the
Tully-Fisher error distribution has non-gaussian wings.
Similarly, the three foreground H I-deficient spirals at
∼ 10 Mpc could have bounced out of the core of the Virgo
cluster, as they lie on the outer edge of the envelope of
the particles after inclusion of 20% rms distance errors in
Figure 3. Indeed, were they bouncing out of the Virgo clus-
ter in the foreground at 15.1 (12.7) Mpc, one would then
have three cases of 34% (21%) errors, each corresponding
to −2.3 σ (−1.3 σ) events for their distance moduli, for
rreb/r100 = 1 (2.5), respectively. Alternatively, they could
lie in the cluster proper, at 16.8 Mpc, which would cor-
respond to three −2.8 σ events for their distance moduli,
which again appears unlikely, unless the Tully-Fisher error
distribution has non-gaussian wings.
The offset of a galaxy relative to the Tolman-Bondi
locus in Figure 2 can be caused by a large distance error
or a large peculiar velocity relative to the general flow
at that distance, or a combination of both. To see which
effect is more important, one can consider the N simulated
particles with distance (with errors folded in) and velocity
respectively within 3 Mpc and 200 km s−1 of each galaxy
in Table 7, as well as angular distance to the halo center
within 1.◦5 of that of the galaxy relative to M87, and asked
what fractions P1 and P2.5 of these particles were located
(before the distance errors were folded in) within 1 or 2.5
times the virial radius of the halo, respectively.
These fractions must be taken with caution, because
the groups falling in or moving out of the Virgo cluster are
at different distances and with different relative masses be-
tween the simulations and the observations. For example,
a true group of galaxies in the background of the Virgo
cluster will end up with too high values of P1 and P2.5 if
there is no group at the same distance in the simulations,
while a Virgo cluster galaxy with a measured distance of
10 Mpc or so beyond the cluster will have too low val-
ues of P1 and P2.5 if the simulations include a group at a
distance close to this measured distance. Nevertheless, P1
and P2.5 should provide interesting first order constraints.
Of course, this method of locating galaxies, implies
that galaxies NGC4064, NGC 4580 and NGC 4586, at
angular distances from M87 beyond θ100 + 1.◦5 = 7.◦1 have
P1 = 0. Interestingly, all three galaxies are likely to be
within 2.5 r100 (i.e. P2 is close to unity).
Among the 5 foreground galaxies with θ < 5.◦6, NGC
4312, NGC 4438, M90 and NGC 4606 are all likely to be in
the Virgo cluster and all very likely to be close enough that
they may have passed through its core. Only NGC 4424,
apparently at distance 4.1 Mpc, is too close to actually be
located within 2.5 virial radii from the Virgo cluster.
Among the 5 background galaxies with θ < 5.◦6, all
but NGC 4305, lying at 40 Mpc, have a non negligible
probability of having crossed the cluster. Only VCC 1569
measured at 23 Mpc is highly likely to be located within
the virial radius of the Virgo cluster. Interestingly, the 3
galaxies at measured distances around 28 Mpc have very
different values of P1 and P2.5. The closest one, NGC 4307,
has one chance in two of being located near the cluster and
40% probability of being within the cluster itself. However,
the furthest one, NGC 4356, has one chance in 5 of belong-
ing to the Virgo cluster and less than one chance in four
of being within 2.5 virial radii. The third one, NGC 4492,
has only one chance in 9 of being within 2.5 virial radii of
the Virgo cluster. This low probability is not surprising,
given that it is located close to the infall/expansion zone
of Figure 2.
12
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
4.2. Are the outlying H I-deficient galaxies located
within small groups?
The alternative to ram pressure stripping of gas is removal
of gas by tidal effects. One can envision various scenarios:
1) tidal stripping of gas beyond the optical radius; 2) tidal
compression of gas clouds leading to gas transformation
into stars; 3) tidal stripping of the gas reservoir infalling
into the disk. Such tides can occur from interacting galax-
ies or from the cluster or group itself. We begin by asking
whether the H I-deficient galaxies apparently outside of
Virgo lie in groups.
Figure 4 shows the environment of the H I-deficient
galaxies in three equal logarithm distance bins, ±1.28 ×
σ ln D wide, centered around 10 Mpc (the apparent dis-
tance to the foreground H I-deficient galaxies), 16.8 Mpc
(the distance to the Virgo cluster) and 28 Mpc (the ap-
parent distance to the background H I-deficient galax-
ies), as well as in two radial velocity bins. The log dis-
tance (velocity) bins are chosen wide enough that if the
parent distribution function of the log distance (veloc-
ity) errors were gaussian, one would have a probability
2 erf(1.28) − 1 = 0.86 of having a galaxy with a true dis-
tance (velocity) within the interval appearing to lie at the
center of the interval.
Among the 5 foreground H I-deficient spirals, 4 (NGC
4312, NGC 4438, and M90, which, as noted in Sect. 3.5,
is far off the Tolman-Bondi relation, and NGC 4606), ap-
pear superposed (within dotted circle in upper-left plot)
with the galaxies at the distance of the main structure of
the Virgo cluster, which suggests that their distances may
be seriously underestimated, unless their happens to be a
group of galaxies precisely along the line of sight to the
center of Virgo lying at 10 Mpc.
One may be tempted to identify the 3 H I-deficient spi-
rals close to M87 (NGC 4438, M90, and NGC 4606) with a
group of ellipticals seen in the upper-left plot of Figure 4.
The velocity dispersion of this group would explain why
these 3 galaxies are so far off the Tolman-Bondi solution
given in Figure 2. However, none of the ellipticals in this
region have accurate (< 38%) relative distance estimates
(cf. Table 2) and some or all may actually belong to the
Virgo cluster.
Interestingly, 3 of the 4 H I-deficient spirals at 28 Mpc
(NGC 4307, NGC 4356 and NGC 4492) lie in a dense re-
gion, extending to the SSW of the position of M87 (lower-
left plot), which is also seen in the middle distance bin
(middle-left plot) corresponding to the distance to M87,
but with 3 times fewer galaxies. Although the eastern edge
of the region, and particularly the H I-deficient NGC 4492,
coincide with the position of M49 (NGC 4472), it appears
distinct from M49, which roughly lies at the middle of a
group of galaxies (middle-left plot) that does not seem to
be associated with the dense region at 28 Mpc (lower-left
plot).
We now compare the Tully-Fisher distance estimator
with the redshift distance estimator Dv = v/H0. If a struc-
ture outside of the Virgo cluster lies at a distance D within
a region with 1D velocity dispersion σv, the error on the
distance estimator Dv = v/H0 will be set by the disper-
sion of the peculiar velocities:
δDv = σv/H0 .
(11)
The redshift distance estimator Dv will be more accurate
than the direct distance estimator when
H0 D >
,
(12)
σv
δD/D
where δD/D is the relative error of the direct distance esti-
mator. For Tully-Fisher distance measurements, δD/D ≃
20% and a field velocity dispersion of 40 − 85 km s−1 (as
observed by Ekholm et al. 2001 and Karachentsev et al.
2003, as well as predicted by Klypin et al. 2003), equa-
tion (12) yields a maximum distance for believing individ-
ual direct distances of 3 to 7 h−1
2/3 Mpc, so that individual
galaxy distances of 28 Mpc are much less reliable.
This change in optimal distance estimators is illus-
trated in Figure 1, where only one of the background H I-
deficient galaxies has a radial velocity much larger than
the mean of the Virgo cluster. The two plots on the right of
Figure 4 show that the structure to the South-Southwest
(SSW) of the Virgo cluster lies close to Virgo in velocity
space (i.e. the concentration is in the right middle plot).
In redshift space, the 3 background H I-deficient galaxies
to the SSW of M87 remain in a similar concentration of
galaxies (middle right plot of Fig. 4) as in distance space
(lower left plot), but this concentration is at similar radial
velocities as the Virgo cluster. However, only one (NGC
4307) has a reasonably high chance (P1 in Table 7) of be-
ing located in or near the Virgo cluster, while one (NGC
4356) has less than one chance in 4 of being in or near
the cluster (from the analysis of Sect. 4.1), and one (NGC
4492) has a radial velocity indicative of a background ob-
ject expanding away from the Virgo cluster (this galaxy
shows up as the empty triangle in Fig. 2 very close to the
Tolman-Bondi locus), and has only roughly one chance in
9 of being even close to the cluster. But the analysis of
P1 and P2.5 of Sect. 4.1 should be superior to the analy-
sis of Fig. 4, because it takes into account the correlation
of distance and velocity. In summary, the 3 background
H I-deficient galaxies to the SSW of M87 may be suffering
from tidal effects of a background group of galaxies.
4.3. Close companions and mergers
According to a search in SIMBAD, we find that (at least) 4
of the 13 H I-deficient galaxies in Table 7 have companions
whose mass and proximity may generate tides that could
remove some of the neutral Hydrogen, either by stripping
the gas beyond the optical radius (without stripping the
stars, thus leaving an H I-deficiency) or by compressing
the diffuse neutral hydrogen, converting atomic gas into
molecular gas, then into stars. Of course, the presence of
a companion in projection, even with a similar velocity,
does not guarantee that this companion is in physical in-
teraction with the galaxy in question, as projection effects
are important in clusters.
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
13
Fig. 4. Positions of galaxies within 9◦ of M87 in 3 distance (left ) and velocity (right ) bins. Ellipticals and spirals are
shown as circles and triangles, respectively. Messier ellipticals are shown as filled circles, while M49 (NGC 4472) is
also highlighted as a thick dotted circle (bottom plots, in green in the Electronic version of the journal). H I-deficient
(> 3 σ) spirals are shown as filled triangles (with the 13 apparently in the outskirts of Virgo [Table 7] in red in the
14
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Table 8. Companions to H I-deficient galaxies in the out-
skirts of Virgo
Galaxy
VCC NGC
522 4305
1043 4438b
1690 4569c
1859 4606
∆v
Companion
VCC NGC (km s−1)
4306
523
1030
4435
1686d --
4607
1868
−380
−214
1354
593
∆mc
T d/ropt M K
0.9a
1.0
4.2
0.6
2.8 R? --
I
1.0
I
1.1
R R
I? --
2.7
Notes: a B-band; b also Arp 120; c also M90 and Arp 76; d also
IC 3583.
Table 8 shows the properties of the companions to
these 4 galaxies. Column 5 is the difference in velocity
relative to the Local Group; column 6 is the difference
in extinction-corrected total I-band magnitude; column 7
is the angular distance between the two components, in
units of the optical radius (D25/2) of the major member;
columns 8 and 9 are indicators for morphological and kine-
matical (Rubin et al. 1999) disturbances (R → regular, I
→ irregular) for the major component. Table 8 indicates
that NGC 4438 has a very close major companion, which
is likely to be responsible for its irregular morphology and
internal kinematics. NGC 4305 and NGC 4606 have mod-
erately close major companions. M90 has a close very mi-
nor (≈ 1/50 in mass) companion.
Moreover, NGC 4307 and NGC 4356 are close in 4D
space: their projected separation (at the distance of M87)
is 215 kpc and their velocity difference is 85 km s−1, so
they may constitute an interacting pair. Similarly, NGC
4580 and NGC 4586 are at a projected separation of 311
kpc with a velocity difference of 254 km s−1, so they too
may constitute an interacting pair.
Finally, two other galaxies in Table 7 have unusually
low inner rotation (Rubin et al. 1999): NGC 4064 and
NGC 4424. Kenney et al. (1996) argue that NGC 4424
has undergone a merger with a galaxy 2 to 10 times less
massive and presumably the same can be said for NGC
4064.
We can obtain some clues about the origin of the H I
deficiency of the outlying galaxies by studying their star
formation rates. We adopt the, obviously simplified, point
of view that tidal interactions are believed to enhance star
formation (e.g. Liu & Kennicutt 1995), while ram pressure
stripping should reduce star formation. We therefore verify
the tidal interaction hypothesis by checking for enhanced
star formation rates. Because tidal interactions operate
on scales of order the orbital times of interacting pairs,
typically 250 Myr or more, we do not use the Hα star
formation rate indicator, as it is only sensitive to very
recent star formation, but focus instead on the broad-band
color B−H. We have taken B and H magnitudes from the
GOLDMine database (Gavazzi et al. 2003).
Figure 5 shows the B−H colors as a function of H-band
luminosity for the 12 NGC galaxies of Table 7 (we have
omitted VCC 1569, which is too underluminous to be in
GOLDMine) and for other Virgo galaxies. The colors and
luminosities in this figure were corrected for absorption
using internal and galactic (Schlegel, Finkbeiner, & Davis
1998) absorption obtained from LEDA, assuming selective
extinction coefficients from Cardelli, Clayton, & Mathis
(1989). Figure 5 clearly displays a general color-luminosity
relation, i.e. the trend that more luminous galaxies are
redder. Comparing galaxies of similar mophological types,
one checks that among the four candidates for possess-
ing tidal companions, NGC 4305, NGC 4438, NGC 4569
(M90), and NGC 4606 (see Table 8), the first three are
blue for their luminosities, assuming their Tully-Fisher
distances, but given that the latter three might well lie
in Virgo (see column P1 in Table 7), we also check that
at the distance to the Virgo cluster these 3 galaxies are
very blue (NGC 4438), extremely blue (M90) and just
blue (NGC 4606). The blue colors for these 4 galaxies,
given their luminosity and morphological type, confirms
the tidal hypothesis for all four objects.
possibly
interacting
the
Interestingly,
galaxies
NGC 4307 and NGC 4356 have blue colors confirming
their interaction if they are indeed in the background
of Virgo as their Tully-Fisher distances suggest, but
have red colors as expected for ram pressure stripped
galaxies if they lie in the Virgo cluster. The values of
P1 in Table 7 cannot distinguish between these two
possibilities. The situation is even more uncertain for the
possibly interacting galaxies NGC 4580 and NGC 4586,
whose Tully-Fisher distances place them slightly behind
and in front of the Virgo cluster, respectively. If they both
lie at the distance of the Virgo cluster, the first one would
have a normal color, while NGC 4586 would be blue as
expected if it is indeed interacting with NGC 4580.
The situation is also confusing for the two galaxies
with possible recent minor merging: NGC 4064 appears
normal or red if it lies in the foreground as suggested by
its Tully-Fisher distance, but blue if it lies at the distance
of the Virgo cluster, and NGC 4424 appears normal or red
if it lies as close as its Tully-Fisher distance suggests but
quite blue if it lies at the distance of Virgo. One would
expect the minor merger to generate a blue nucleus if the
merging took place recently enough, but this may have
little effect on the overall color of the galaxy.
Moreover, one may contest that tides will remove stars
as well as gas from the disk of a spiral galaxy. However, the
estimates of H I-deficiency presented here are based upon
single-dish (Arecibo) radio observations, which often do
not have sufficient angular resolution to map the gas (the
half-power beam of Arecibo subtends 3.′3, close to the typ-
ical optical angular diameters of the H I-deficient galaxies
-- see Table 7). One can therefore easily imagine that
fairly weak tidal effects will deplete the neutral Hydrogen
gas beyond the optical radius without affecting the stars
in the spiral disk, thus leading to an H I-deficiency when
normalized relative to the size of the optical disk.
Two galaxies among our H I-deficient Virgo outliers
listed in Table 7 (NGC 4438 and M90) have been studied
by Cayatte et al. (1994), who computed H I-deficiencies
of Virgo spirals, based upon high resolution 21cm VLA
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
15
Fig. 5. Color-luminosity diagram for the 12 of the 13 H I-deficient galaxies of Table 7 (VCC 1569 was excluded because
it is not bright enough to be included in GOLDMine), possibly outside the Virgo cluster (large filled symbols) or placed
at the distance of the Virgo cluster (large open symbols). Other Virgo galaxies are shown as small open symbols.
Triangles, squares and pentagons (respectively in red, green, and blue in the Electronic version of the journal) refer to
galaxies of morphological type 1− 2, 3− 4 and 5− 6, respectively. The data is taken from the GOLDMine (Gavazzi et al.
2003) database, and dereddended using the internal and Galactic extinctions from LEDA with the selective extinctions
of Cardelli, Clayton, & Mathis (1989).
observations. Both galaxies present very small 21cm to
optical diameter ratios, indicating that the gas deficiency
sets in at small radii and cannot be caused by tidal effects.
Additional high-resolution 21 cm observations at the VLA
of the other galaxies in Table 7 will obviously allow con-
firmation of this result.
4.4. Are the outlying H I-deficient galaxies truly
gas-deficient?
The H I-deficiency estimator has been defined by several
authors (e.g. Chamaraux et al. 1980; Haynes & Giovanelli
1984; Solanes et al. 1996) as the hydrogen mass relative to
the mean for galaxies of the same morphological type and
optical diameter:
DEF1 = hlog MHI(Dopt, T )i − log MHI ,
(13)
16
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
where MHI is the H I mass of the galaxy in solar units.
Solanes et al. (2002) applied a different H I-deficiency es-
timator based on the hybrid mean surface brightness ex-
pected for each galaxy type independently of its diameter,
which has the advantage of being independent of the dis-
tance to the galaxies:
DEF2 = hlog ΣHI(T )i − log ΣHI ,
where ΣHI is the hybrid mean surface brightness, defined
as
(14)
ΣHI =
FHI
θ2
opt
,
(15)
where FHI and θopt are the H I flux and optical angular
diameter, respectively.
Could the seemingly H I-deficient galaxies in the fore-
ground and background of Virgo have normal gas content
according to equation (13) but appear deficient with equa-
tion (14)? Solanes et al. checked the statistical agreement
of these two H I-deficiency estimators by fitting a linear
model between the two parameters and found a regres-
sion line with slope and intercept close to the unit and
zero respectively, with a scatter or 0.126. Figure 6 shows
this correlation, where those galaxies possible located out-
side the Virgo cluster are shown as solid circles. The plot
clearly indicates that the outliers are also deficient with
the DEF1 definition of H I-deficiency.
Fig. 6. Comparison of H I-deficiency estimators. Galaxies
possibly located outside the Virgo cluster are shown as
filled circles. The vertical and horizontal dashed lines
shows the 3 σ cutoff for DEF1 and DEF2, respectively.
However, Table 7 clearly shows that 7 out of our
13 H I-deficient possible outliers appear to be Sa galax-
ies (T = 1), some of which might be misclassified
S0/a or even S0 galaxies, which are known to have
little cold gas. The stronger H I-deficiency of early-
type spirals in the Virgo cluster has been already no-
ticed by Guiderdoni & Rocca-Volmerange (1985). Some
of these Sa galaxies may not be deficient relative to
field S0/a or S0 galaxies. Indeed, the most recent study
(Bettoni, Galletta, & Garc´ıa-Burillo 2003) of the H I con-
tent of galaxies as a function of morphological type in-
dicates that the logarithm of the H I mass normalized to
luminosity or to square optical diameter decreases increas-
ingly faster for earlier galaxy types, therefore its slope for
Sa's is considerably larger than for Sc's, and hence S0/a
galaxies misclassified as Sa's will appear more H I-deficient
than later-type galaxies misclassified by one-half of a mor-
phological type within the Hubble sequence.
Moreover,
independent estimates of morphological
types by experts led to σ(T ) = 1.5 (rms) for galaxies with
optical diameters > 2′ (Naim et al. 1995). The galaxies
listed in Table 7 have optical diameters ranging from 2′
to 10′ plus one (VCC 1569) with D = 0.′8. There are
two galaxies (NGC 4305 and NGC 4492) with T = 1 and
D ≤ 2′, i.e. where the uncertainty on T is probably as
high as 1.5.
For example, NGC 4424 and NGC 4492 are marginally
H I-deficient and are prime candidates for losing their sta-
tus of gas-deficient if they are misclassified lenticulars (by
∆T = 1, the typical scatter given by LEDA).
We must thus resort to visualizing the possibly mis-
classified galaxies. Figure 7 displays the snapshots of the
7 H I-deficient outliers classified as Sa (T = 1). We now
discuss the morphologies of these 7 galaxies.
NGC 4064 This galaxy has a very large high surface
brightness inner disk or flattened bulge with little trace
of spiral arms. NGC 4064 may lie between S0 and S0/a.
NGC 4305 This galaxy is of type Sa or later. It has a
nearby major companion.
NGC 4424 This galaxy has an important bar and the
size of the bulge is difficult to estimate (Kenney et al.
1996 fit the bulge-disk ratio to between 0.2 and 0.6). It
shows little traces of spiral arms. Kenney et al. write
"NGC 4424 is classified as an Sa rather than an S0
because the outer stellar disk is not structureless". It
may well be an S0/a.
NGC 4438 This galaxy, with a large bulge, is tidally
distorted by a companion just North of the frame.
Gavazzi et al. (2000) fit a pure r1/4 law to this galaxy,
suggesting that its morphology is of an earlier type
than Sa.
NGC 4492 This galaxy is viewed face-on, hence the disk
is not very visible. The bulge is relatively large and
their are signs of a weak spiral arm. NGC 4492 may
lie between S0/a and Sa, but Gavazzi et al. (2000) fit
a bulge to disk ratio of 1/3, suggesting a type of Sa or
later.
NGC 4586 This galaxy shows spiral arms and a normal
bulge for an Sa galaxy.
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
17
NGC 4064
NGC 4305
NGC 4424
NGC 4438
NGC 4492
NGC 4586
NGC 4606
Fig. 7. POSS I snapshots of the T = 1 H I-deficient galaxies of Table 7. The box size is 6′ and North points upwards.
NGC 4606 This galaxy shows little traces of spiral arms
and has an intermediate bulge. NGC 4606 appears be-
tween S0/a and Sa. However, Gavazzi et al. (2000) fit
a bulge to disk ratio of 1/20, suggesting a later type
than Sa.
Altogether, we suspect that 2 galaxies may be earlier than
Sa: NGC 4064 and NGC 4424, both of which appear to
have recently ingested smaller galaxies, and significant
amounts of cold gas may have been heated up during the
merger.
Given that between types S0/a and Sa, the compilation
of Bettoni et al. (2003) (left plots of their Fig. 2) indicates
that the normalized H I content varies as log M norm
∼
T /4 + cst, an overestimate of 1 (0.5) in T will lead to an
underestimate in H I-deficiency of 0.25 (0.125). The H I-
deficiency of NGC 4064 is sufficiently high that a correc-
tion of 0.25 will still keep it among the 3 σ cases using both
DEF1 and DEF2 criteria. On the other hand, if the mor-
phological type of NGC 4424 is S0/a (T = 0), it would
still be H I-deficient with criterion DEF1, but no longer
among the 3 σ cases with DEF2.
HI
18
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Note, however, that if most H I-deficient outliers clas-
sified as Sa's were in fact S0/a's, we would find ourselves
with the new problem of justifying the presence of S0/a's
so far away from the cluster core and apparently not
within groups (see section 5.3). We also have to keep in
mind that Koopmann & Kenney (1998) argued that spi-
rals in the Virgo cluster could have also been misidentified
as early-type objects, an opposite trend to the above sug-
gested.
4.5. Case by case analysis
We now focus on each of the 13 galaxies, one by one.
NGC 4064 The
very
rotation
(Rubin, Waterman, & Kenney 1999) suggests that it
is a merger remnant (see NGC 4424 below), which
may have been responsible for its H I-deficiency.
inner
low
NGC 4305 This galaxy, at 40 Mpc and with a large ra-
dial velocity, is almost certainly a background galaxy.
It has a major tidal companion (Table 8) and a blue
color for its luminosity, assuming it is indeed far behind
Virgo, confirming the tidal origin of its gas deficiency.
NGC 4307 The origin of the H I-deficiency in this galaxy
is difficult to ascertain, as it has 41% (48%) probability
of lying within the sphere centered on the Virgo clus-
ter with radius its 1 (2.5) times the virial radius. Its
proximity with NGC 4356 suggests that it may have
interacted in the past with that galaxy, and if it is
indeed in the background of Virgo, as suggested by
its Tully-Fisher distance, it would be blue for its lu-
minosity, confirming the interaction hypothesis with
NGC 4356.
NGC 4312 This galaxy is very likely within the Virgo
cluster.
NGC 4356 The origin of the H I-deficiency in this galaxy
is difficult to ascertain, as it has 19% (23%) probability
of lying within the sphere centered on the Virgo clus-
ter with radius its 1 (2.5) times the virial radius. Its
proximity with NGC 4307 suggests that it may have
interacted in the past with that galaxy, as confirmed
by its blue color for its luminosity, assuming it lies in-
deed in the background of Virgo.
NGC 4438 This
NGC 4424 The very low inner rotation of this galaxy
suggests a merger remnant (Rubin et al. 1999). It al-
most certainly lies in the foreground of the Virgo clus-
ter. Moreover, as discussed in Sect. 4.4, NGC 4424 may
be of type S0/a, in which case it would not be 3 σ de-
ficient in neutral Hydrogen.
seemingly
is
the Virgo cluster.
very probably a member of
Kotanyi & Ekers
the H I-
deficiency of this highly distorted galaxy is caused not
by the tidal companion, which should increase the
line-width, but by interaction with the intracluster
medium of the Virgo cluster. Moreover, this galaxy
has a very small radial extent of
its H I gas, as
attested by the VLA map of Cayatte et al. (1994).
foreground
suggest
galaxy
(1983)
that
Interestingly, at the distance of Virgo, this galaxy
would be very blue, which is easily explained by the
presence of its tidal companion.
NGC 4492 This galaxy is probably in the background
of the Virgo cluster, in which case it would be blue for
its luminosity and morphological type. It may lie in a
group that could include NGC 4307 and NGC 4356 and
thus be suffering tidal encounters with these and/or
other members of the group.
VCC 1569 This faint galaxy probably lies in or very
close to the Virgo cluster.
is
still
the optical
rotation curve
M90 Among the 13 galaxies listed in Table 7, this
very blue Sab LINER (e.g. Keel 1996) deviates
the most significantly from both the Tolman-Bondi
locus (Fig. 2) and from the color-luminosity di-
agram (Fig. 5).
Its distance has been controver-
sial. Stauffer, Kenney, & Young (1986) suggest that
the Tully-Fisher estimate is an underestimate be-
cause
rising
its measurement (see Rubin et al.
to the limit of
1999) and that similarly low distances derived by
Cowley, Crampton, & McClure (1982) using their line
index / bulge luminosity relation may be erroneous be-
cause the weakness of the nuclear lines is in contrast
with the high metallicity expected from the stellar pop-
ulation of the bulge.
M90 is extremely isolated in projected phase space.
Among the galaxies with D < 13 Mpc, it appears
associated with a group of ellipticals, none of which
has accurate distance measurements, with sky coordi-
nates associated with the central ellipticals with accu-
rate distance measurements in the center of the Virgo
cluster. Given its bright corrected total blue magni-
tude of 9.63 (Table 7), it's absolute magnitude would
correspond to MB = −20.24 at 9.4 Mpc or MB =
−21.5 at 16.8 Mpc. The latter absolute magnitude is
very bright for a spiral galaxy. This can be quanti-
fied by comparing with the turnover luminosity, L∗,
of the field galaxy luminosity function. Unfortunately,
the surveys used to derive the field galaxy luminos-
ity function do not employ the Johnson B band used
here. However, the 2MASS survey has measured a to-
tal K-band magnitude of K = 6.58 for M90. For a
turnover absolute magnitude (Kochanek et al. 2001)
M ∗
K = −23.39 + 5 log h = −24.27 for h = 2/3 adopted
here, the K-band luminosity of M90 corresponds to
0.4 L∗ if M90 is at 9.4 Mpc but 1.3 L∗ if it lies at the
distance of the Virgo cluster, which is a large lumi-
nosity for a spiral galaxy. Moreover, at the distance of
Virgo, M90 would be extremely blue for its luminosity
(Fig. 5).
Finally, the contours of continuum radio emission seen
from the NVSS survey taken at the VLA are com-
pressed in the East-Northeast direction, which is op-
posite to the direction towards the center of the Virgo
cluster. This suggests that M90 may be bouncing out
of the cluster, which would also explain the galaxy's
negative radial velocity.
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
19
NGC 4580 This galaxy probably lies near but not in the
Virgo cluster, and may possibly be bouncing out of the
cluster, therefore losing its neutral gas by ram pressure
stripping. It may be interacting with NGC 4586, and
is somewhat blue for its luminosity, especially if it lies
behind Virgo.
NGC 4586 This galaxy probably lies near the Virgo
cluster and may possibly be bouncing out of the clus-
ter, therefore losing its neutral gas by ram pressure
stripping. It may be interacting with NGC 4580, and
is blue for its luminosity, especially if it lies within the
Virgo cluster.
NGC 4606 This galaxy is a very likely member of the
Virgo cluster. It is likely to be tidally perturbed by its
companion NGC 4607 (Table 8), as attested by its blue
color for its luminosity assuming it indeed lies in Virgo,
although it shows only modest signs of such tides.
4.6. Origin of the H I-deficiency in the outlying galaxies
The H I-deficient galaxies apparently lying well in front
and behind of the Virgo cluster must have passed through
the core of the cluster to have shed their gas by stripping
by the ram pressure of the intracluster gas. As shown by
Mamon et al. (2004), galaxies cannot bounce out beyond
1 to 2.5 virial radii, which according to their estimate of
the virial radius of Virgo, amounts to 1.7 to 4.1 Mpc.
Therefore, given a distance to the Virgo cluster of 16.8
Mpc, we suggest that only galaxies with distances between
12.7 and 20.9 Mpc could have passed through Virgo.
One possibility, discussed in Sect. 4.1, is that the true
3D locations of the outlying H I-deficient galaxies are in
fact within the Virgo cluster's virial radius or at most
within 4 Mpc from its center, which is possible (although
not likely) given the typical 20% relative distance errors
for the spiral galaxies in Virgo, expected from the Tully-
Fisher relation, and from our comparison of the observed
velocity distance relation with that derived from cosmo-
logical N -body simulations. Indeed, of the 13 H I-deficient
spirals appearing in the outskirts of the Virgo cluster, 4
to 5 (depending on how to place the distance to M90)
are highly likely to be located in 3D within the virial ra-
dius of the Virgo cluster (Table 7) and one other galaxy
(NGC 4307) has one chance in two of being in the cluster
or near it. Three others are at an angular radius > θ100
from M87, but appear to be within 2.5 virial radii of it
in 3D. Finally, two galaxies (4356 and NGC 4492) in the
direction of the cluster have low but non negligible proba-
bility of being in or near the cluster, while the last two are
definitely foreground (NGC 4424) and background (NGC
4305) galaxies. Hence, between 4 and 11 of these 13 galax-
ies may have passed through the core of the Virgo cluster
and seen their neutral gas ram pressure swept by the hot
intracluster gas.
The presence of a few discrepant distance measure-
ments suggests that the nearly gaussian parent distri-
bution function for measurement errors in the distance
modulus (or relative distance) has extended non-gaussian
wings.
It is also possible that highly inaccurate (at the 30%
level or more) Tully-Fisher distance measurements may
arise from physical biases. If galaxies are H I-deficient be-
cause of strong interactions with the cluster (through ram
pressure stripping or tides) or with neighboring galax-
ies (through collisional tides), their Tully-Fisher distances
may be biased and/or inefficient. Indeed, our analysis of
Sect. 4.3 indicates that, among the galaxies likely or possi-
bly within the Virgo cluster, one (NGC 4438) shows clear
tidal perturbations from a close major companion, while
2 others (M90 and NGC 4606) are probably tidally per-
turbed by close companions, and the one possible Virgo
cluster member (NGC 4307) may have interacted with a
fairly close major companion in the past.
Moreover, there are general arguments, independent of
these case-study tidal effects, that suggest that abnormal
Tully-Fisher distances may be caused by H I-deficiency.
For example, Stauffer et al. (1986) noted that H I-deficient
spirals in the direction of the Virgo cluster tend to have
smaller line-widths for their H-band luminosity, and, sim-
ilarly, Rubin, Hunter, & Ford (1991) found that the max-
imum rotation velocity of spiral galaxies within com-
pact groups of galaxies tend to be smaller for their lu-
minosity than for normal spirals. Such small line-widths
or maximum rotation velocities imply that distances are
underestimated. Using 2D (Fabry-Perot) spectroscopy,
Mendes de Oliveira et al. (2003) have refined the analysis
of Rubin et al. and find that the lowest luminosity galax-
ies in their sample (MB = −19.5±1) have lower maximum
rotation velocities than field spirals of the same luminosity.
Moreover, they find that compact group galaxies display
a more scattered Tully-Fisher relation than that of field
spirals.
For galaxies lying in or very near the Virgo cluster,
an alternative to ram pressure gas stripping is for tidally
stripped gaseous reservoirs that prevent subsequent gas in-
fall to the disk, which quenches subsequent star formation
(Larson, Tinsley, & Caldwell 1980). However, if a spiral
galaxy lies in a cluster, it may well be on its first infall,
provided that the effective ram pressure stripping of its
neutral gas quickly converts it into an S0 galaxy. It seems
reasonable to expect that tidally stripped gas reservoirs
will not have time to have a large effect on the gas content,
so that such galaxies with tidally stripped gas reservoirs
will not be as severely H I-deficient as the 13 3 σ cases of
Table 7.
The H I-deficient galaxies that are truly outside of the
cluster may have lost their gas by the tides generated dur-
ing close encounters with other galaxies within a group
or small cluster of galaxies. Indeed, the efficiency of tidal
perturbations of galaxies from close encounters with other
galaxies is stronger in groups and small clusters than in
rich clusters (see Fig. 5 of Mamon 2000), basically because
the rapid motions in rich clusters decrease the efficiency of
tides. Three of the four H I-deficient galaxies at 28 Mpc lie
within a significant concentration of galaxies (Figs. 1 and
20
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
4) that ought to create such collisional tides, even with
the estimates of distance using the more precise radial ve-
locities (for these distant objects, see eq. [12]).
The 3 galaxies outside the virial angular radius (NGC
4064, NGC4580 and NGC4586) all probably lie closer than
2.5 virial radii from M87, and may be in the process of
bouncing out of the cluster after having lost their H I gas
through ram pressure stripping by the intracluster gas, if
the maximum rebound radius is closer to the upper limit of
2.5 virial radii found by Mamon et al. (2004), than of their
lower limit of 1 virial radius. One of these galaxies (NGC
4064) is a candidate for a recent merger as witnessed by
its very low inner rotation, while the other two (NGC
4580 and NGC 4586) may have interacted in the past,
given their proximity in phase space. In the case of recent
merging, the neutral gas may have been shock heated with
that of the galaxy being swallowed up, and should shine
in Hα or X-rays depending on its temperature.
Of the 4 galaxies likely or certainly far outside the
Virgo cluster (in 3D), one (NGC 4305) may be tidally per-
turbed by a close major companion, one (NGC 4424) has
suffered a recent merger (given its very low inner rotation
velocity), one may have interacted with another galaxy
(NGC 4307) in the past, and one (NGC 4492) may be a
misclassified S0/a (with T = 0), implying a lower expected
H I content, hence a less than 3 σ H I-deficiency upon the
estimator DEF1 and just at 3 σ upon DEF2, although the
bulge/disk ratio derived by Gavazzi et al. (2000) suggests
that the galaxy is indeed an Sa.
Although we can provide at least one explanation for
each case of H I-deficiency in the 13 spiral galaxies that
1) appear to be located in the foreground or background
of the Virgo cluster or 2) lie just outside the projected
virial radius of the cluster, there is certainly room for an
improved analysis. In particular, accurate distances ob-
tained with Cepheids would be most beneficial, as well as
VLA maps for 11 of the 13 galaxies to distinguish tidal
stripping from ram pressure stripping.
Acknowledgements. We wish to thank Philippe Amram,
Hector Bravo-Alfaro, V´eronique Cayatte, Avishai Dekel, Peppo
Gavazzi, Gopal Krishna, and Gilles Theureau for useful discus-
sions, and an anonymous referee for very useful comments. We
also thank Fran¸cois Bouchet, Bruno Guiderdoni and cowork-
ers for kindly providing us with their N -body simulations, and
Jeremy Blaizot for answering our technical questions about the
design and access to the simulations. TS acknowledges hospi-
tality of the Institut d'Astrophysique de Paris where most of
this work was done, and she and GAM acknowledge Ewa Lokas
for hosting them at the CAMK in Warsaw, where part of this
work was also done. TS was supported by a fellowship of the
Ministerio de Educaci´on, Cultura y Deporte of Spain. We have
used the HyperLEDA database (http://leda.univ-lyon1.fr)
operated at the Observatoire de Lyon, France, SIMBAD
(http://simbad.u-strasbg.fr), operated by the CDS in
Strasbourg, France, the NASA/IPAC Extragalactic Database
(NED, http://nedwww.iap.caltech.edu), which is operated
by the Jet Propulsion Laboratory, California Institute of
Technology, under contract with the National Aeronautics and
Space Administration, and the GOLDMine database operated
by the Universita' degli Studi di Milano-Bicocca.
Appendix A: Median and half-width of the
likelihood function for a set of measurements
with different errors
For a single measurement of variable X of value x with
uncertainty σi, assuming gaussian errors, the probability
of measuring xi is
1
exp(cid:20)−
(xi − x)2
(cid:21) .
2 σ2
i
(A.1)
√2πσi
pi(xix) =
The likelihood of measuring a set of {xi} given a true value
x is
L({xi}x) = Yi
= Yi
pi(xix)
(cid:18) 1
√2πσi(cid:19) exp(cid:18)−
2 (cid:19) ,
(A.2)
S(x)
with
S(x) = x2 Xi
1
σ2
i − 2 xXi
xi
σ2
i
+Xi
x2
i
σ2
i
.
(A.3)
From Bayes' theorem, the probability of having a true
value x is
f (x)
(A.4)
g({xi}) L ({xi}x) ,
P (x{xi}) =
where f (x) is the a priori probability of having a true
distance x and g({xi}) is the a priori probability of mea-
suring the set of {xi}. If the measurements are made with
a selection function near unity, then g ≃ 1.
For simplicity, one can assume a uniform f (x). The
probability of the true X being less than some value x is
then
P (< x) = R x
−∞ L({xi}x) dx
R ∞
−∞ L({xi}x) dx
= R x
−∞ exp(cid:2)−x2/2 Pi 1/σ2
R ∞
−∞ exp [−x2/2 Pi 1/σ2
erfc(cid:18) ν
√2(cid:19) ,
1
2
=
i + x Pi xi/σ2
i + x Pi xi/σ2
i(cid:3) dx
i ] dx
(A.5)
where
i
i
.
(A.6)
that the median X is
i − x Pi 1/σ2
pPi 1/σ2
ν = Pi xi/σ2
Now erfc(ν/√2)/2 = 0.16, 0.5, 0.84 for ν = −1, 0, +1, so
x50 = Pi xi/σ2
Pi 1/σ2
while the width equivalent to the standard deviation of a
single gaussian is
(A.7)
,
i
i
σ =
1
2
(x84 − x16) =
1
pPi 1/σ2
i
.
(A.8)
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
21
Since L is an exponential function of x (eq. [A.2]), the
mean and mode will be equal to x50.
Alternatively, one could adopt an underlying physical
model for f (x). If X is the distance modulus, for which the
gaussian p.d.f. of equation (A.1) should be adequate, one
could use a Virgocentric model, where the probability f (x)
stems from a density distribution around Virgo, using for
example an NFW model. If the angle between the galaxy
and Virgo, as seen by the observer, is θ, one has
r2 = D2
10 {dex(−0.4 d) + dex(−0.4 dV )
−2 cos θ dex[−0.2 (d + dV )]} ,
(A.9)
where d and dV are the distance moduli to the galaxy and
to Virgo, respectively, and D10 = 10 pc. Hence, one has
,
dr
dd
dr2
dd
f (d) ∝ ρ(r)
1
dr
2 r
dd
0.2 ln 10 D2
10
=
=
r
(A.10)
× [cos θ dex(−0.2(d + dV )) − dex(−0.4d)] .(A.11)
The problem with this approach is that the underly-
ing model, i.e. ρ(r), is uncertain: Should one take a single
NFW model at the location of M87? Should one make
ρ(r) uniform at a radius where ρNFW(r) reaches the mean
mass density of the Universe? Or better, should one re-
place this uniform background with a collection of other
halos, as expected in the hierarchical scenario? Moreover,
because the model requires a distance to M87, one would
need to iterate using a first guess on the distance to M87
(perhaps based upon the estimate with a uniform f (x)).
These questions led us to assume in this paper a uniform
distribution for f (x).
References
Bettoni, D., Galletta, G., & Garc´ıa-Burillo, S. 2003, A&A,
405, 5
Bevington, P. R. & Robinson, D. K. 1992, Data reduction
and error analysis for the physical sciences (New York:
McGraw-Hill, 2nd ed.)
Ekholm, T., Baryshev, Y., Teerikorpi, P., Hanski, M. O.,
& Paturel, G. 2001, A&A, 368, L17
Ekholm, T., Lanoix, P., Teerikorpi, P., Fouqu´e, P., &
Paturel, G. 2000, A&A, 355, 835
Ferguson, H. C. & Binggeli, B. 1994, A&A Rev., 6, 67
Gavazzi, G., Boselli, A., Donati, A., Franzetti, P., &
Scodeggio, M. 2003, A&A, 400, 451
Gavazzi, G., Boselli, A., Scodeggio, M., Pierini, D., &
Belsole, E. 1999, MNRAS, 304, 595
Gavazzi, G., Franzetti, P., Scodeggio, M., Boselli, A., &
Pierini, D. 2000, A&A, 361, 863
Giovanelli, R. & Haynes, M. P. 1985, ApJ, 292, 404
Guiderdoni, B. & Rocca-Volmerange, B. 1985, A&A, 151,
108
Gunn, J. E. & Gott, J. R. 1972, ApJ, 176, 1
Hatton, S., Devriendt, J., Ninin, S., et al. 2003, MNRAS,
343, 75
Haynes, M. P. & Giovanelli, R. 1984, AJ, 89, 758
-- . 1986, ApJ, 306, 466
Jensen, J. B., Tonry, J. L., & Luppino, G. A. 1998, ApJ,
505, 111
Karachentsev, I. D., Makarov, D. I., Sharina, M. E., et al.
2003, A&A, 398, 479
Keel, W. C. 1996, PASP, 108, 917
Kenney, J. D. P., Koopmann, R. A., Rubin, V. C., &
Young, J. S. 1996, AJ, 111, 152
Klypin, A., Hoffman, Y., Kravtsov, A. V., & Gottlober,
S. 2003, ApJ, 596, 19
Kochanek, C. S., Pahre, M. A., Falco, E. E., et al. 2001,
ApJ, 560, 566
Koopmann, R. A. & Kenney, J. D. P. 1998, ApJ, 497, L75
Kotanyi, C. G. & Ekers, R. D. 1983, A&A, 122, 267
Kundu, A. & Whitmore, B. C. 2001, AJ, 121, 2950
Larson, R. B., Tinsley, B. M., & Caldwell, C. N. 1980,
ApJ, 237, 692
Liu, C. T. & Kennicutt, R. C. 1995, ApJ, 450, 547
Mamon, G. A. 2000,
in 15th IAP Astrophys. Mtg.,
Dynamics of Galaxies:
from the Early Universe to
the Present, ed. F. Combes, G. A. Mamon, &
V. Charmandaris, Vol. 197 (San Francisco: ASP), 377 --
388, astro-ph/9911333
Mamon, G. A., Sanchis, T., Salvador-Sol´e, E., & Solanes,
J. M. 2004, A&A, in press, astro-ph/0310709
Binggeli, B., Sandage, A., & Tammann, G. A. 1985, AJ,
Mendes de Oliveira, C., Amram, P., Plana, H., &
90, 1681
Balkowski, C. 2003, AJ, 126, in press
Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ,
Naim, A., Lahav, O., Buta, R. J., et al. 1995, MNRAS,
345, 245
274, 1107
Cayatte, V., Kotanyi, C., Balkowski, C., & van Gorkom,
J. H. 1994, AJ, 107, 1003
Cayatte, V., van Gorkom, J. H., Balkowski, C., & Kotanyi,
Neilsen, E. H. & Tsvetanov, Z. I. 2000, ApJ, 536, 255
Ninin, S. 1999, PhD thesis, Universit´e de Paris XI
Rubin, V. C., Hunter, D. A., & Ford, W. Kent, J. 1991,
C. 1990, AJ, 100, 604
ApJS, 76, 153
Chamaraux, P., Balkowski, C., & G´erard, E. 1980, A&A,
Rubin, V. C., Waterman, A. H., & Kenney, J. D. P. 1999,
83, 38
AJ, 118, 236
Cole, S. & Lacey, C. 1996, MNRAS, 281, 716
Cowley, A. P., Crampton, D., & McClure, R. D. 1982,
Sanchis, T., Solanes, J. M., Salvador-Sol´e, E., Fouqu´e, P.,
& Manrique, A. 2002, ApJ, 580, 164
ApJ, 263, 1
Sandage, A., Binggeli, B., & Tammann, G. A. 1985, AJ,
Davis, M., Efstathiou, G., Frenk, C. S., & White, S. D. M.
90, 1759
1985, ApJ, 292, 371
22
T. Sanchis et al.: H I-deficiency on outskirts of Virgo. II
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ,
500, 525
Solanes, J. M., Giovanelli, R., & Haynes, M. P. 1996, ApJ,
461, 609
Solanes, J. M., Manrique, A., Garc´ıa-G´omez, C., et al.
2001, ApJ, 548, 97
Solanes, J. M., Sanchis, T., Salvador-Sol´e, E., Giovanelli,
R., & Haynes, M. P. 2002, AJ, 124, 2440
Spergel, D. N., Verde, L., Peiris, H. V., et al. 2003, ApJS,
148, 175
Stauffer, J. R., Kenney, J. D., & Young, J. S. 1986, AJ,
91, 1286
Teerikorpi, P., Bottinelli, L., Gouguenheim, L., & Paturel,
G. 1992, A&A, 260, 17
Tonry, J. L., Dressler, A., Blakeslee, J. P., et al. 2001,
ApJ, 546, 681
Tully, R. B. & Fisher, J. R. 1977, A&A, 54, 661
Tully, R. B. & Shaya, E. J. 1984, ApJ, 281, 31
Young, C. K. & Currie, M. J. 1995, MNRAS, 273, 1141
|
astro-ph/0207365 | 1 | 0207 | 2002-07-17T17:14:24 | Studying the Gaseous Phases of Galaxies with Background QSOs | [
"astro-ph"
] | High resolution rest frame UV quasar absorption spectra covering low and high ionization species, as well as the Lyman series lines, provide remarkably detailed information about the gaseous phases of galaxies and their environments. For redshifts less than 1.5, many important chemical transitions remain in the observed ultraviolet wavelength range. I present examples of absorption that arises from lines of sight through a variety of structures, drawn from UV spectra recently obtained with STIS/HST. Even with the greater sensitivity of COS/HST there will be a limit to how many systems can be studied in detail. However, there is a great variety in the morphology of the phases of gas that we observe, even passing through different regions of the same galaxy. In order to compile a fair sample of the gaseous structures present during every epoch of cosmic history, hundreds of systems must be sampled. Multiple lines of sight through the same structures are needed, as well as some probing nearby structures whose luminous hosts have been studied with more standard techniques. Combined with high resolution optical and near--IR ground--based spectra, it will be possible to uniformly study the gaseous morphologies of galaxies of all types through their entire evolutionary histories. | astro-ph | astro-ph |
Hubble's Science Legacy: Future Optical-Ultraviolet Astronomy from Space
ASP Conference Series, Vol. 000, 2002
K.R. Sembach, J.C. Blades, R.C. Kennicutt, G.D. Illingworth (eds.)
Studying the Gaseous Phases of Galaxies with
Background QSOs
Jane C. Charlton
Department of Astronomy & Astrophysics, The Pennsylvania State
University, 525 Davey Laboratory, University Park, PA 16802
Abstract. High resolution rest frame UV quasar absorption spectra
covering low and high ionization species, as well as the Lyman series
lines, provide remarkably detailed information about the gaseous phases
of galaxies and their environments. For redshifts less than 1.5, many
important chemical transitions remain in the observed ultraviolet wave-
length range. I present examples of absorption that arises from lines of
sight through a variety of structures, drawn from UV spectra recently
obtained with STIS/HST. Even with the greater sensitivity of COS/HST
there will be a limit to how many systems can be studied in detail. How-
ever, there is a great variety in the morphology of the phases of gas that
we observe, even passing through different regions of the same galaxy.
In order to compile a fair sample of the gaseous structures present dur-
ing every epoch of cosmic history, hundreds of systems must be sampled.
Multiple lines of sight through the same structures are needed, as well as
some probing nearby structures whose luminous hosts have been studied
with more standard techniques. Combined with high resolution optical
and near -- IR ground -- based spectra, it will be possible to uniformly study
the gaseous morphologies of galaxies of all types through their entire evo-
lutionary histories.
1.
Introduction
The tool of quasar absorption line spectroscopy has several distinct advantages
over more traditional imaging studies of distant galaxies. Spectra covering ab-
sorption lines from numerous chemical elements in various states of ionization
can yield detailed information about the physical conditions in the gaseous com-
ponents of the universe. This is not limited to only the most luminous compo-
nents. Dwarf galaxies, low surface brightness galaxies, and even intergalactic
structures are probed by quasar lines of sight. Quasar absorption lines can be
used to study gas during the birth and death of stars and of entire galaxies. The
kinematic information contained in high resolution spectra allows us to study
processes. This is not just a still picture snapshot; it is more like a short movie.
Finally, the same level of detail is available for our study at all redshifts because
the same method can be applied using optical and near -- IR spectroscopy.
However, the study of quasar absorption lines also presents some challenges
if we are to extract from the quasar spectra all of the detailed information about
1
2
Charlton
physical conditions of gas. Imagine if we had to classify a galaxy according to
its standard morphological type by zooming in on an image of just a small part
of the galaxy. We have to consider carefully what a single line of sight tells us
about the global conditions in galaxies. In fact, it should be possible to learn
a great deal if we consider the evolution of the ensemble of gaseous structures
probed by quasar lines of sight. In order to realize this potential, we must learn
to connect absorption signatures to the physical conditions of the phases of gas
that produce them, and to the processes that give rise to such signatures in local
galaxies.
For the optimal study of the detailed physical conditions along a quasar
line of sight we need high resolution spectroscopy covering all rest frame UV
transitions. For redshifts less than one, a significant fraction of the key transi-
tions still appear in the ultraviolet. In this proceedings, I review examples of
various systems that have been studied in detail, focusing on the question of
what additional data would yield a significant advance. The goal is to define
capabilities for the optimal UV spectrograph and telescope to be used for such
a program.
2. A C iv -- Deficient Strong Mg ii Absorber?
The z = 0.99 absorber toward the quasar PG1634+706 is a strong Mg ii absorber,
which implies that it is very likely to be within an impact parameter of 35h−1 kpc
of an ∼ L∗ galaxy (Steidel 1995). It is part of the small subset of Mg ii absorbers
that have particularly small C iv absorption features, i.e. it is C iv -- deficient rela-
tive to other strong Mg ii absorbers (Churchill et al. 2000). STIS/E230M spectra
of this quasar have been obtained by Jannuzi and by Burles, and HIRES/Keck
spectra (covering Mg i, Mg ii, and Fe ii ) have been obtained by Churchill and
Vogt (2001). The most important transitions are shown in Figure 1. Detailed
results of our modeling of this system have been presented by Ding et al. (2002).
The system can be described by a combination of a minimum of four differ-
ent phases of gas. A phase is a region or regions with density and temperature
within some range, spatially separate from other phases. The Mg ii absorption
arises in clouds that have densities of ∼ 0.01 cm−3. The bulk of the Mg i ab-
sorption cannot arise in the same clouds. Instead, we propose that the Mg i is
produced by small (∼ 100 AU), cool pockets of the interstellar medium. The
metallicity of the Mg ii and Mg i clouds is about 0.2×solar. A "broad", low-
metallicity (< 0.01×solar) component is required to self -- consistently fit Ly α,
Ly β, and the higher -- order Lyman -- series lines. The low metallicity could relate
to why this absorber is C iv deficient. Finally, a strong, smooth Si iv absorption
profile suggests an additional collisionally ionized phase with T ∼ 60, 000 K,
perhaps heated by shocks.
For this system, there are two outstanding issues of particular interest.
First, with higher resolution (R > 100, 000) we could test the narrow Mg i cloud
hypothesis. Second, an image of the host galaxy (or those for similar systems)
would enable us to see whether an early -- type galaxy is responsible for the lack
of absorption resembling the Milky Way corona.
Gaseous Phases of Galaxies
3
1
0
1
0
1
0
1
0
1
0
1
0
-300
-200
-100
0
100
200
300
Figure 1.
Selected transitions for the z = 0.9902 strong Mg ii system
along the PG 1634 + 706 line of sight, presented in velocity space. The
velocity zero -- point corresponds to the apparent optical depth centroid
for the Mg ii profile. The Mg i and Mg ii profiles were obtained with
HIRES/Keck, with R = 45, 000 (P.I. Churchill), and the other transi-
tions were observed with STIS/HST, with R = 30, 000 (P.I.'s Jannuzi
and Burles).
4
Charlton
3. A Redshift One Galaxy Group?
The z ∼ 0.93 systems in the PG 1206 + 459 line of sight are the subject of
a separate contribution in this volume by Jie Ding, and therefore no detail or
figures will be presented here. Three distinct systems (two strong Mg ii and
one weak Mg ii) are apparent at z = 0.925, z = 0.928, and z = 0.934, an
overall velocity spread of just over 1000 km s−1. This is likely to be a group of
galaxies, with the three systems arising in three very different types of galaxy
hosts. The study of these systems would be aided by higher S/N since there
are suggestions of abundance pattern variations. Understanding of the phases
giving rise to high ionization absorption would be much better constrained if a
high resolution spectrum covering the O vi transition was obtained.
4. A Pair of Dwarf Galaxies?
The z = 1.04 Absorber Toward PG 1634 + 706 is a multiple cloud, weak Mg ii
absorber, with two sets (kinematic subsystems) of low ionization "clouds" sep-
arated by ∼ 150 km s−1. Each of the subsystems has an O vi absorption profile
offset by ∼ 50 km s−1. Profiles for key transitions from the STIS/HST and
HIRES/Keck spectra are shown in Figure 2. This system was modeled by Zonak
et al. (2002), and the results are summarized here.
The Mg ii clouds have metallicity < 0.03×solar, constrained by a partial
Lyman limit break. Although Si iv has the same kinematic structure as Mg ii in
the stronger subsystem, the clouds appear to be offset by velocities ranging from
3 -- 5 km s−1. This could be related to an H ii region flow, but the interpretation
of the multiple cloud structure remains unclear. The Si iii profile is smooth, but
unsaturated, and is stronger than can be explained by the combination of the
Mg ii and Si iv cloud phases. This suggests a collisional ionization mechanism
for producing Si iii. The Ly α profile requires two additional photoionized phases
that can also produce the broad O vi profiles.
The low metallicity and the kinematics of the two subsystems suggest that
two dwarf galaxies and their halos could be responsible. The lower ionization
phases would arise in the inner regions of the dwarfs, and the offset higher
ionization phases from their halos.
Very little is known about the expected absorption signature of dwarf galax-
ies. If quasar absorption line probes through nearby dwarfs were available, this
mystery would be quickly solved. However, this would require that some fainter
quasars were accessible for high resolution spectroscopy.
5. A Variety of Weak Mg ii Absorbers
Figure 3 shows selected transitions for the three single -- cloud, weak Mg ii ab-
sorbers at z = 0.65, z = 0.81, and z = 0.90 along the same quasar PG 1634+706
line of sight. Unlike the strong Mg ii absorbers, single -- cloud, weak Mg ii ab-
sorbers are generally not found within 50 kpc of L∗ galaxies. Rigby, Charlton,
& Churchill (2002) found, based on a larger sample for which lower resolution
FOS spectra were available, that they cannot be far below solar metallicity.
They also found that at least a subset of these absorbers are very small in di-
Gaseous Phases of Galaxies
5
1
0
1
0
1
0
1
0
1
0
1
0
1
0
-300
-200
-100
0
100
200
300
Velocity [km/s]
Figure 2.
Selected transitions for the z = 1.0414 multiple -- cloud,
weak Mg ii system along the PG 1634 + 706 line of sight, presented in
velocity space. These are taken from the same spectra noted in the
Figure 1 caption.
6
Charlton
Figure 3.
A comparison of selected transitions for the three single --
cloud, weak Mg ii absorbers along the PG1634 + 706 line of sight. The
spectra from which these were extracted are referenced in the Figure 1
caption.
Gaseous Phases of Galaxies
7
mension. Detailed models of the three PG 1634 + 706 single -- cloud, weak Mg ii
absorbers were presented in Charlton et al. (2002).
The z = 0.81 weak Mg ii absorber is the kinematically simplest along this
line of sight. The C iv absorption is relatively weak. It was not even detected in
a low resolution FOS spectrum. The Lyα profile constrains the metallicity to be
supersolar, and the ionization conditions indicate that the low ionization phase
is sub -- parsec scale. Despite the C iv being weak, the broad shape of its profile
cannot be fit by the Mg ii cloud. The high ionization phase could be collisionally
ionized, but photoionization is more likely, with a cloud size of ∼ 100 pc. The
high ionization phase could surround or be surrounded by the low ionization
gas.
If the high ionization material is a thicker shell surrounding a small low
ionization pocket then we would expect to see many systems with only the high
ionization phase. A large statistical sample is needed to determine the geometry.
To distinguish between collisional ionization and photoionization, coverage of the
O vi transition is key.
In contrast, the z = 0.90 weak Mg ii absorber has relatively strong C iv.
A low ionization phase has supersolar metallicity and a cloud size of ∼ 50 pc.
The C iv and N v require a separate, broader cloud, which is consistent with
photoionization and not with collisional ionization. Also, in this case, an addi-
tional offset high ionization cloud is required to fit the blue wing of the C iv and
the Ly α, which is not centered on the Mg ii. The C iv equivalent width could
be large in this case partly because the line of sight passed through a separate
cloud at a different velocity, and not because the centered C iv cloud has special
physical conditions.
The C iv profile associated with the z = 0.65 weak Mg ii absorber has dis-
tinct subcomponent structure, and therefore a large equivalent width. The Lyα
is also quite strong for this system, possibly also because of the large kinematic
spread. It is centered on the C iv and not on the single Mg ii component. The low
ionization phase is poorly constrained, but its metallicity must exceed 0.1×solar.
The three C iv components arise in high ionization, kiloparsec -- scale structures,
which are likely to be photoionized.
What are these weak, single -- cloud Mg ii absorbers? The small size con-
straints on the low ionization phases implies that there must be huge numbers
of the host structures.
In fact, to account for the observed number of weak,
single -- cloud Mg ii absorbers, per unit redshift, there must be more than a mil-
lion hosts for every L∗ galaxy in the universe. Yet, these are not associated with
L∗ galaxies. They appear to be self -- enriched pockets of higher density material
embedded in higher ionization, more diffuse gas. Rigby et al. (2002) discuss the
origin of the weak Mg ii absorbers at length. Their two most likely explanations
are small traces of gas in fading remnants of Population III star clusters and
fragments of supernova shells in faint blue dwarf galaxies.
A number of new capabilities would help to constrain the properties of
this intriguing class of object. Double line of sight observations at a scale of
parsecs, though they could not be done for many systems, would provide an
essential check on size constraints for these weak Mg ii absorbers. Nearby weak
Mg ii absorbers must be identified so that narrow band imaging can be used
to discover if hosts are dwarf galaxies or isolated star clusters. Of course, a
8
Charlton
large statistical sample would clarify whether there are different classes of weak,
single -- cloud Mg ii absorbers, with different types of hosts.
6. Conclusions
The small sample of systems presented here demonstrates great promise for
learning about the detailed physical conditions in a variety of gaseous environ-
ments. For these systems, the data have some limitations and, as a result,
some questions remain unanswered. These highlight the need for additional UV
spectroscopy capabilities. I conclude by presenting a "wish list" if we aim to
construct the dynamical history of the ensemble of galaxies and gaseous struc-
tures:
• hundreds of systems, or even thousands!
• cover all key chemical species from the rest frame ultraviolet
• study many systems that are at low enough redshift to image their galaxy
hosts; these serve as a calibrator for higher redshift systems
• higher spectral resolution (R = 100, 000 or even higher) for a subset of the
systems in order to resolve interstellar medium features
• high signal -- to -- noise (> 20) for a subset to enable abundance pattern stud-
ies
• multiple lines of sight through the same objects
• a separate, detailed study of the interstellar medium of the Milky Way and
of nearby galaxies using the same techniques
Support for this work was provided by the NSF (AST -- 9617185) and by
NASA (NAG 5 -- 6399 and HST -- GO -- 08672.01 -- A), the latter from the Space Tele-
scope Science Institute, which is operated by AURA, Inc., under NASA contract
NAS5 -- 26555.
References
Charlton, J. C., Ding, J., Zonak, S. G., Churchill, C. W., & Bond, N. A. 2002,
ApJ, submitted
Churchill, C. W., Mellon, R. R., Charlton, J. C., Jannuzi, B. T., Kirhakos, S.,
& Steidel, C. C., 2000, ApJ, 543, 577
Churchill, C. W., & Vogt, S. S. 2001, AJ, 122, 679
Ding, J., Charlton, J. C., Bond, N. A., Zonak, S. G., & Churchill, C. W. 2002,
ApJ, submitted
Ding, J., Charlton, J. C., Churchill, C. W., & Palma, C. 2002, in preparation
Rigby, J. R., Charlton, J. C., & Churchill, C. W. 2002, ApJ, 565, 743
Steidel, C. C. 1995, in QSO Absorption Lines, ed. G. Meylan (Garching:Springer --
Verlag), 139
Zonak, S. G., Charlton, J. C., Ding, J., & Churchill, C. W. 2002, ApJ, submitted
|
0711.4549 | 1 | 0711 | 2007-11-28T18:04:46 | Hydrogen Isocyanide in Comet 73P/Schwassmann-Wachmann (Fragment B) | [
"astro-ph"
] | We present a sensitive 3-sigma upper limit of 1.1% for the HNC/HCN abundance ratio in comet 73P/Schwassmann-Wachmann (Fragment B), obtained on May 10-11, 2006 using Caltech Submillimeter Observatory (CSO). This limit is a factor of ~7 lower than the values measured previously in moderately active comets at 1 AU from the Sun. Comet 73P/Schwassmann-Wachmann was depleted in most volatile species, except of HCN. The low HNC/HCN ratio thus argues against HNC production from polymers produced from HCN. However, thermal degradation of macromolecules, or polymers, produced from ammonia and carbon compounds, such as acetylene, methane, or ethane appears a plausible explanation for the observed variations of the HNC/HCN ratio in moderately active comets, including the very low ratio in comet 73P/Schwassmann-Wachmann reported here. Similar polymers have been invoked previously to explain anomalous 14N/15N ratios measured in cometary CN. | astro-ph | astro-ph |
To appear in The Astrophysical Journal
Preprint typeset using LATEX style emulateapj v. 03/07/07
HYDROGEN ISOCYANIDE IN COMET 73P/SCHWASSMANN-WACHMANN (FRAGMENT B)
D.C. Lis1,4, D. Bockel´ee-Morvan2, J. Boissier2, J. Crovisier2, N. Biver2, and S.B. Charnley3
(Received 4 May 2007; Accepted 27 November 2007)
To appear in The Astrophysical Journal
ABSTRACT
We present a sensitive 3-σ upper limit of 1.1% for the HNC/HCN abundance ratio in comet
73P/Schwassmann-Wachmann (Fragment B), obtained on May 10 -- 11, 2006 using Caltech Submil-
limeter Observatory (CSO). This limit is a factor of ∼7 lower than the values measured previously in
moderately active comets at 1 AU from the Sun. Comet 73P/Schwassmann-Wachmann was depleted
in most volatile species, except of HCN. The low HNC/HCN ratio thus argues against HNC production
from polymers produced from HCN. However, thermal degradation of macromolecules, or polymers,
produced from ammonia and carbon compounds, such as acetylene, methane, or ethane appears a
plausible explanation for the observed variations of the HNC/HCN ratio in moderately active comets,
including the very low ratio in comet 73P/Schwassmann-Wachmann reported here. Similar polymers
have been invoked previously to explain anomalous 14N/15N ratios measured in cometary CN.
Subject headings: comets: individual (73P/Schwassmann-Wachmann) -- molecular processes -- radio
lines: solar system
1. INTRODUCTION
Hydrogen isocyanide, HNC, a metastable isomer of
HCN, was first detected in the ISM over a third of a cen-
tury ago as the unidentified line U90.7 (Snyder & Buhl
1972; Zuckerman et al. 1972; the identification subse-
quently confirmed in the laboratory by Blackman et al.
1976). The interstellar HNC/HCN abundance ratio
has been shown to be strongly temperature dependent
(e.g., Schilke et al. 1992; Hirota et al. 1998, and ref-
erences therein). The current understanding of the
HNC chemistry in molecular hot cores is that in warm,
dense gas HCN is first produced efficiently by ion-
molecule chemistry from ammonia and C+. HNC is
subsequently produced by proton transfer to HCN to
form HCNH+, followed by dissociative recombination or
proton transfer to ammonia (e.g., Rodgers & Charnley
2001a; Charnley et al. 2002). The temperature depen-
dence of the HNC/HCN abundance ratio is explained
primarily by proton transfer reactions cycling between
the two isomers via the HCNH+ ion. However, ad-
ditional HNC formation routes may be required in
dark clouds where the HNC/HCN ratio exceeds unity.
These are presumably neutral-neutral reactions (e.g.,
C+NH2=HNC+H; Herbst, Terzieva, & Talbi 2000) and
consequently are probably too slow to efficiently form
HNC in expanding cometary atmospheres.
The initial detection of HNC in a cometary atmosphere
was by Irvine et al. (1996) in comet C/1996 B2 (Hyaku-
take). The measured HNC/HCN abundance ratio, ∼6%,
was similar to that in interstellar clouds with a temper-
ature of order 50 K, suggesting that cometary HNC may
1 California Institute of Technology, Downs Laboratory of
Physics 320-47, Pasadena, CA 91125; [email protected]
2 LESIA and UMR 8109 du CNRS, Observatoire de
Paris, 5 place Jules Janssen, 92195 Meudon, France; Do-
[email protected],
[email protected],
[email protected]; [email protected]
3 Space Science
and Astrobiology Division, MS 245 -- 3,
NASA Ames Research Center, Moffet Field, CA 94035; charn-
[email protected]
4 Visiting Scientist, LESIA, Observatoire de Paris
be unprocessed interstellar material incorporated into
the comet's nucleus. However, Irvine et al.
(1996) ar-
gued that a number of alternative processes may also ex-
plain the observed HNC/HCN ratio in comet Hyakutake,
including irradiation of icy matrix containing HCN, non-
equilibrium chemical processes in the solar nebula, gas-
phase processes in the coma itself, infrared relaxation of
HCN from excited vibrational levels of the ground elec-
tronic state, or photo-dissociation of a heavier parent
molecule.
A strong variation of the HNC/HCN abundance ra-
tio in comet C/1995 O1 (Hale-Bopp) with heliocentric
distance (from ∼<2% at 2.9 AU to ∼20% near 1 AU;
Biver et al. 1997; Irvine et al.
1998) questioned the
interstellar origin of cometary HNC and suggested a
production mechanism in the coma itself as a more
likely explanation. Rodgers & Charnley (1998) pre-
sented a comprehensive model of the cometary coma
chemistry and suggested that in very active comets,
such as comet Hale-Bopp, the observed variation of the
HNC abundance with the heliocentric distance, as ob-
served with single-dish telescopes, can be explained by
isomerization of HCN driven by the impact of fast hy-
drogen atoms produced in photo-dissociation of par-
ent molecules. However, their model overproduced the
HNC abundance at ∼3 AU by about a factor of 2.
In a subsequent paper Rodgers & Charnley (2001b)
showed that the same mechanism cannot reproduce ob-
served HNC/HCN abundance ratios in moderately active
comets at ∼1 AU, such as comet C/1999 H1 (Lee). The
applicability of the model to very active comets has also
been questioned by interferometric observations of HNC
and HCN in comet Hale-Bopp (Bockel´ee-Morvan et al.
2005; Bockel´ee-Morvan 2007), which show that the
HNC/HCN visibility ratio is almost flat as a function
of baseline, and equal to the single-dish flux ratio. This
suggests that, contrary to model predictions, HNC in
comet Hale-Bopp was produced efficiently in the inner
coma (parent scale-length ≪ 2000 km).
Comet 73P/Schwassmann-Wachmann is a short-period
2
Lis et al.
(5.34 years) comet of the Jupiter family that likely origi-
nated from the trans-Neptunian population in the outer
Solar System. Narrow-band photometry and spectro-
scopic observations during the 1990 apparition showed
that this comet is strongly depleted in C2 and NH2 rel-
ative to CN and OH, placing it in the 21P/Giacobini-
Zinner class of odd comets with extreme depletions
(A'Hearn et al. 1995; Fink & Hicks 1996). During its
1995 perihelion passage, comet 73P split into several
fragments (Boehnhardt et al. 2002).
Its 2006 appari-
tion thus offered an exceptional opportunity to measure
the HNC/HCN ratio in fragments of a Jupiter-family
comet of deviant composition. Indeed, comet 73P made
a very close approach to Earth in May 2006 (at 0.067
and 0.079 AU for Fragments B and C, respectively), that
compensated for its moderate activity (water production
rate of a few × 1028 s−1 at perihelion) and allowing sen-
sitive investigations.
In the present paper, we compile the existing mea-
surements of the HNC/HCN ratio in moderately active
comets, including our sensitive upper limit obtained in
comet 73P/Schwassmann-Wachmann and discuss possi-
ble implications for the origin of HNC in cometary at-
mospheres.
2. OBSERVATIONS
Observations of Comet 73P/Schwassmann-Wachmann
(Fragment B) presented here were carried out on May 9 --
11, 2006 UT using the 10.4-m Leighton telescope of Cal-
tech Submillimeter Observatory (CSO) on Mauna Kea
in Hawaii. The heliocentric and geocentric distances of
the comet at the time of the CSO observations were 1.02
and 0.072 AU, respectively. We used the 345 GHz facil-
ity SIS receiver and acousto-optical spectrometers with
total bandwidths of 50 MHz and 1 GHz. The frequency
scale of the spectrometers (the reference channel and the
channel width) was established by injecting calibration
signals from 10 MHz and 100 MHz frequency comb gen-
erators, for the 50 MHz and 1 GHz AOS, respectively.
The effective frequency resolution of the high-resolution
(50 MHz bandwidth) spectrometer was 95 kHz (2 chan-
nels), corresponding to ∼0.08 km s−1. Pointing of the
telescope, checked by performing five-point continuum
scans of Jupiter in the south, was determined to drift
by ∼<5′′ on time-scales of ∼3 hours at the time of the
observations (one fifth of the ∼25′′ FWHM CSO beam
at 360 GHz). The accuracy of the cometary ephemeris,
checked by performing five-point scans of the HCN J=4 --
3 transition in the comet, was determined to be ∼7 -- 9′′.
The main beam efficiency, determined from total-power
observations of Jupiter, Saturn and Mars, was measured
to be ∼74%. The absolute calibration uncertainty of the
individual measurements is ∼20%. However, since the
HCN and HNC lines were observed almost simultane-
ously, with the same receiver, we estimate the calibration
uncertainty of the HNC/HCN line ratio to be smaller, of
order 15%.
HCN J=4 -- 3 spectra (rest frequency 354.50547 GHz)
taken on May 9 -- 11 UT are shown in Figure 1 and the
line intensities are listed in Table 1, along with the cor-
responding statistical uncertainties, as determined from
the noise level measured in the spectra. A significant de-
crease in the HCN intensity (by a factor of 1.8) is seen
over the period of our observations, following an earlier
Fig. 1. -- HCN J=4 -- 3 spectra in comet 73P/Schwassmann-
Wachmann (Fragment B) obtained on May 9 -- 11 (left to right, re-
spectively). Gray curves in the two right panels show the May 9
spectrum.
Fig. 2. -- HNC J=4 -- 3 spectra in comet 73P/Schwassmann-
Wachmann obtained on May 10 -- 11 (left and middle panels, re-
spectively). The right panel shows an uniformly weighted average
of the two data sets. Gray curves show HCN J=4 -- 3 spectra on the
corresponding days, scaled down by a factor of 50.
outburst. The low-level pedestal, seen most clearly in
the May 9 spectrum, is due to the HCN hyperfine struc-
ture (e.g., Lis et al. 1997). HNC J=4 -- 3 spectra (rest
frequency 362.63031 GHz) taken on May 10 -- 11 UT are
shown in Figure 2. The HNC and HCN observations on
these two nights were interleaved in time to account for
the time variability of the outgassing. No HNC emission
is detected (Table 1).
Fragment B of comet 73P underwent several outbursts
during the April -- May 2006 apparition, shedding tens of
fragments, as imaged for example by the Hubble Space
Telescope around 18 April (e.g., Weaver et al. 2006).
On 8.0 May UT fragment B entered a new outburst
phase, beginning as a strong brightness increase of its
central condensation. Visual magnitudes reported in
International Comet Quarterly (2006) showed a fivefold
increase in total brightness the following day and a 5 to
10 times increase in the water outgassing rate was seen
in the Nan¸cay and Odin data between 7 and 10 of May
(Crovisier et al. 2006; Colom et al. 2006).
HNC in Comets
3
TABLE 1
HCN and HNC Line Intensities.
U T Date
May 9, 2006
May 10, 2006
May 11, 2006
May 10 -- 11, 2006
I(HCN)
I(HNC)
--
5.35 ± 0.04
0.073 ± 0.025
4.21 ± 0.08
3.01 ± 0.06 −0.021 ± 0.018
3.62 ± 0.05
0.026 ± 0.015
Adv (K km s−1), computed
Notes. -- Integrated line intensities, T ∗
over the velocity range of -- 2 to 2 km s−1 (HCN; includes all hy-
perfine components) and -- 1 to 1 km s−1 (HNC). The uncertainties
listed are 1-σ statistical uncertainties, as determined from the noise
level measured in the spectra. The May 10 -- 11 entry is a uniformly
weighted average of the two corresponding spectra.
Molecular production rates of HCN and HNC (3-σ up-
per limits) on May 10 -- 11 are listed in Table 2. The
resulting 3-σ upper limit for the HNC/HCN abundance
ratio, based on the average of the May 10 and May 11
data is 1.1%.
IRAM 30-m observations impose upper
limits of 2.9 and 3.8% for the HNC/HCN ratio on May
14 and 18, respectively (Biver et al. 2006a), consistent
with the low value presented here.
3. HNC/HCN RATIO IN COMETS
HNC has now been detected in 11 comets with wa-
ter production rates in the range 3 × 1028 − 4 ×
1030 mol s−1 (e.g., Irvine et al. 1996; Biver et al. 2005;
Crovisier et al. 2005; Biver et al. 2006b), not count-
ing the very active comets Hale-Bopp and C/2006 P1
(McNaught), where different chemical processes may be
at work. The observed HNC/HCN abundance ratio is
plotted in Figure 3 as a function of the water produc-
tion rate. The plot appears to show a distinction be-
tween comets with water production rates above and be-
low ∼4 × 1029 mol s−1. More productive comets seem to
display a relatively high HNC/HCN ratio, ∼0.2, with lit-
tle scatter. The scatter increases significantly in weaker
comets and no clear dependence of the observed abun-
dance ratio on the water production rate is seen. Since
water production rates in comets vary with the helio-
centric distance and the measurements included in our
sample span a wide range in rh, we have outlined with
gray circles in Figure 3 the measurements with rh lim-
ited to a narrow range of 0.5 -- 0.8 AU. This sub-sample
includes four comets, which cover a wide range of water
production rates. No dependence of the HNC/HCN ratio
on the water production rate is seen. Therefore we con-
clude that, as explained below, the apparent distinction
between the more productive and weak comets in Fig-
ure 3 is simply caused by selection effects, as all active
comets in our sample happened to be observed at small
heliocentric distances.
Figure 4 shows the dependence of the HNC/HCN ratio
on the heliocentric distance of the comet. A clear vari-
ation is seen and, and for rh between ∼0.5 and 1.2 AU,
the ratio is well described by the linear relation
QHNC/QHCN = 0.26 − 0.19 rh(AU) ,
(1)
shown as the dotted line in Figure 4. The formal 1σ
uncertainties of the intercept and the slope are 0.024
and 0.021, respectively. The two measurements at the
smallest heliocentric distances, rh∼0.1 -- 0.2 AU (comets
C/2002 X5 and C/2002 V1, both with relatively high
Fig.
3. -- Observed HNC/HCN abundance ratio in comets
as a function of the water production rate, QH2O. Measure-
ments at the heliocentric distances between 0.5 and 0.8 AU are
surrounded by gray circles. References: C/1996 B2 (Hyaku-
take):
Irvine et al. 1996; C/1999 H1 (Lee): Biver et al. 2000;
C/1999 S4 (LINEAR): Bockel´ee-Morvan et al.
2001; C/1999
T1 (McNaught-Hartley): Biver et al.
2006b; C/2000 WM1
(LINEAR): Biver et al. 2006b, Irvine et al. 2003; C/2001 A2
(LINEAR): Biver et al. 2006b; 153P/Ikeya -- Zhang: Biver et al.
2006b, Irvine et al. 2003; C/2001 Q4 (NEAT), C/2002 T7 (LIN-
EAR), C/2002 V1 (NEAT), C/2002 X5 (Kudo-Fujikawa), and
C/2004 Q2 (Machholz): Biver et al. (2005), Crovisier et al. 2005;
21P/Giacobini -- Zinner: Biver et al.
1999; 73P/Schwassmann --
Wachmann: this work. The combined uncertainties of the indi-
vidual measurements in the figure are assumed to be equal to the
statistical uncertainties, as measured from the individual spectra,
plus a 15% calibration uncertainty, added in quadrature.
Fig. 4. -- Observed HNC/HCN abundance ratio in comets as a
function of the heliocentric distance, rh. The dotted line shows a
linear fit to the data for rh between ∼0.5 and 1.2 AU. The fitted
slope is 0.192 ± 0.021; the observed variation of the HNC/HCN
abundance ratio with rh is thus statistically significant.
water production rates of ∼4 × 1030 mol s−1), do fall
slightly below the best fit line. This may indicate that
the dependence of the HNC/HCN abundance ratio on
the heliocentric distance flattens for rh∼<0.5 AU. The
least squares fit given above is based on all measure-
ments in the 0.5 -- 1.2 AU heliocentric distance range.
However, a similar rh dependence is obtained from mea-
surements of comets with the water production rates be-
4
Lis et al.
TABLE 2
HCN and HNC Production Rates.
U T Date
May 10, 2006
May 11, 2006
May 10 -- 11, 2006
Q(HCN)
Q(HNC)
Q(HCN)/Q(HNC)
(4.85 ± 0.09) × 1025 < 0.7 × 1024
(3.05 ± 0.06) × 1025 < 0.5 × 1024
(3.90 ± 0.05) × 1025 < 0.4 × 1024
< 1.4%
< 1.7%
< 1.1%
Notes. -- Productions rates (mol s−1) computed assuming a gas temperature of 65 K and expansion velocities of 0.62 and 0.58 km s−1, on
May 10 and 11, respectively. Average pointing offsets of 3′′ and 4′′ for HCN and HNC, respectively, have been assumed. Since the HCN
and HNC transitions studied here have similar rest frequencies and line strengths, the derived Q(HNC)/Q(HCN) ratio is insensitive to the
detailed model assumptions.
low 4 × 1029 mol s−1, which restricts the rh range to
0.75 -- 1.2 AU (the slope given by the least squares fit is
0.144±0.031 vs. 0.191±0.021 above; a ∼1.3σ difference).
The more productive comets in our sample thus follow
essentially the same HNC/HCN dependence on the he-
liocentric distance, as derived for weak comets. The ap-
parent distinction between the two groups in Figure 3 is
caused by the fact that all more productive comets in
our sample happened to be observed preferentially close
to the Sun.5 The observed variation in the HNC/HCN
abundance ratio with the heliocentric distance is likely
related to the change in the temperature of the grains in
the coma, as the comet approaches the Sun.
The 3-σ upper limit of 1.1% for the HNC/HCN ra-
tio in comet 73P/Schwassmann-Wachmann based on our
observations is ∼7 times lower than the typical value
at 1 AU given by the least-squares fit to the measure-
ments in other comets (dotted line in Fig. 4). Given
the small geocentric distance of comet 73P at the time
of our observations (0.072 AU), the HNC production
rate could be significantly underestimated if the emis-
sion originates from an extended source in the coma.
The production rates given above are computed assum-
ing a parent density distribution for both molecules (di-
rect release from the nucleus). The HNC production
rate would increase by a factor of 7 if a daughter den-
sity distribution with a parent scale-length of ∼5000 km
is assumed. The resulting HNC/HCN abundance ra-
tio in comet 73P would then be consistent with other
measurements in moderately active long-period comets
at 1 AU. Observations of comet Hale-Bopp with the
Plateau de Bure Interferometer imply an HNC parent
scale-length ≪ 2000 km (Bockel´ee-Morvan et al. 2005;
Bockel´ee-Morvan 2007). The HNC parent scale-length
has never been constrained directly by interferometric
observations in moderately productive comets. How-
ever, observations of comet 153P/Ikeya-Zhang presented
by Biver et al.
(2006b) suggest that the HNC parent
scale-length in this comet was less than 2500 km (equiv-
alent to less then 7500 km when scaled to 1 AU), with
the best fit Lp = 0. To further investigate the effects
5 Another way to separate the dependence of the HNC/HCN
abundance ratio on rh and QH2O is by fitting the data with a
hQc, with a, b, and c being
combined power-law in the form R = arb
free parameters. The power-law fit diverges at small heliocentric
distances, but in the range of rh considered here, 0.5 -- 1.2 AU, is
a good representation of the measurements. For rh in AU, and Q
in the units of 1029 s−1, the best fit parameters are a = 0.059 ±
0.004, b = −2.30 ± 0.21, and c = −0.09 ± 0.08. This confirms the
presence of a strong dependence of the HNC/HCN ratio on the
heliocentric distance and no statistically significant trend with the
water production rate.
Fig. 5. -- Deviation of the HNC/HCN abundance ratio from the
best fit line described in Fig. 4, in comets with rh between 0.5 and
1.2 AU, plotted as a function of the telescope field of view (in the
units of 5000 km). The dotted line shows a formal least-squares
fit to the data. The fitted slope is 0.0025 ± 0.0063; no statistically
significant trend is seen.
of a possible extended source of HNC on the measured
HNC/HCN abundance ratios, we plot in Figure 5 the
HNC/HCN ratio in the long-period comets observed to
date as a function of the telescope field of view (the lin-
ear trend with the heliocentric distance given by eq. (1)
has been subtracted out). The formal linear fit is shown
as a dotted line; no statistically significant trend is seen.
A combined power-law fit to the HNC/HCN abundance
ratio, as a function of the heliocentric distance and the
field of view, Φ (in the units of 5000 km), in the form
hΦc, gives a = 0.058 ± 0.004, b = −2.03 ± 0.21,
R = arb
and c = 0.10 ± 0.12. This confirms that no trend with
the field of view is apparent in the data. We note that
other comets in our sample, including comet Hyakutake,
were observed with a telescope field of view ∼2000 km.
If the HNC parent scale-length at 1 AU is indeed of or-
der 5000 km, the HNC production rates and the result-
ing HNC/HCN abundance ratios in these objects would
scale up accordingly. While only future sensitive interfer-
ometric observations will be able to unequivocally con-
strain the HNC parent scale-length, a possible extended
source of HNC cannot explain the discrepancy between
the HNC/HCN ratio in comet 73P and the values mea-
sured in other moderately active comets at 1 AU.
Comet 73P/Schwassmann-Wachmann is
the first
short-period (Jupiter-family) comet in which a sensi-
tive upper limit for the HNC/HCN ratio has been
obtained. Our observations might thus indicate that
HNC in Comets
5
HNC is depleted in Jupiter family comets,
for ex-
ample due to evolutionary effects in the ices in the
outer layers of the nucleus. However, comet 73P was
in the process of break-up and fragmentation, with
fresh unprocessed material exposed for the first time
to solar radiation.
Infrared and radio observations
(Dello Russo et al. 2007; Biver et al. 2006a) show re-
markable similarity in the composition of fragments B
and C. All measured volatiles, except HCN, were de-
pleted with respect to water in both fragments, compared
to other comets. Furthermore, no evidence of temporal
variations in the compositions of the fragments during
and in-between outbursts was seen. The explanation of
the low HNC/HCN ratio in comet 73P/Schwassmann-
Wachmann as due to evolutionary effects in the surface
layers of the nucleus thus appears unlikely.
A comparison of the HNC abundance in comet
73P/Schwassmann-Wachmann with that measured in
comet C/1999 S4 (LINEAR) is particularly interesting,
as both comets were in the process of break-up, with
fresh material exposed to the solar radiation. Based
on their IR measurements, Dello Russo et al.
(2007)
argue that relative abundances of parent volatiles in
comet 73P/Schwassmann-Wachmann (with the excep-
tion of HCN) resemble those in C/1999 S4 and are in the
depleted range with respect to the majority of comets
characterized to date (see also Villanueva et al. 2006).
HNC was detected in C/1999 S4 with an abundance
of ∼17% with respect to HCN at a heliocentric dis-
tance of 0.77 AU in a small field of view of 2600 km
(Bockel´ee-Morvan et al. 2001). Correcting for the dif-
ference in the heliocentric distance, as given by eq. (1),
this is a factor of ∼11 higher than the 3-σ upper limit
in 73P/Schwassmann-Wachmann. The HNC abundance
relative to water is also higher in C/1999 S4, by a fac-
tor of ∼8 -- 9; 1.7 × 10−4 (Bockel´ee-Morvan et al. 2001)
vs. ∼<2 × 10−5 (Biver et al. in prep.) We next consider
coma chemistry versus gas-grain processes as means of
producing HNC.
4. ORIGIN OF COMETARY HNC
Rodgers & Charnley (2001b) considered a wide range
of possible explanations for the origin of HNC in mod-
erately active comets. They ruled out the production
by ion-molecule chemistry or by reactions of energetic
hydrogen atoms with HCN. They also ruled out photo-
dissociation of small organic molecules, such as CH2NH,
HNCO, and other N-bearing compounds. This left
large organic molecules or polymers as possible parents
of HNC. One possible candidate was hexamethylenete-
tramine (HMT; C6H12N4), produced by UV irradiation
of simple ices containing ammonia, water, formaldehyde
and methanol. However this compound was shown to
be stable to thermal degradation (Fray 2004) and there-
fore an unlikely parent of HNC. Rodgers & Charnley
(2001b) also suggested HCN polymers (or oligomers),
produced from irradiated HCN molecules in the nuclear
or pre-cometary ices, as a possible parent of HNC (once
HCN tetrameters are formed, they can easily polymerize
to polyaminocyanomethylene (PACM;
[-(NH2)C(CN)-
]n), which can then undergo ring closure to form lad-
der polymers). If such polymers are present in cometary
nuclei, one would expect their abundance in the coma
to be correlated with that of HCN. The abundance
of HCN in comet 73P/Schwassmann-Wachmann im-
plied by IR and radio observations (Biver et al. 2006a;
Villanueva et al. 2006; Dello Russo et al. 2007) is near
the high end of measurements in ∼30 comets observed
at radio wavelengths (Biver et al. 2002, 2006b). PACM
polymers should thus be quite abundant in this par-
ticular comet and the low HNC abundance implied by
our measurements appears to argue against PACM as a
source of cometary HNC.
In cold, dense interstellar material,
A different class of HCN polymers, formed not from
HCN molecules, but instead from ammonia and acety-
lene, has been invoked to explain the low 14N/15N
ratio observed in CN in several comets, a factor of
two below that measured in HCN in comet Hale-Bopp
(e.g., Jehin et al. 2004).
In this scenario, some CN
is produced from photo-dissociation of HCN, however
a secondary, presumably polymeric, source of CN is
needed.
large
15N enhancements may be present in ammonia ices
(Charnley & Rodgers 2002). These ices could be pre-
cursors of cometary material. Further energetic process-
ing is required to incorporate this nitrogen reservoir into
organic material (e.g., addition of NH2 side-groups to
PAH molecules; Charnley 2002). The secondary source
of CN in cometary comae may thus be a polymer pro-
duced from ammonia, since only interstellar ammonia
can obtain sufficiently low 14N/15N ratios, and from car-
bon compounds, such as acetylene, methane, or ethane.
The initial analysis of the Stardust organic data
(Sandford et al. 2006) indicates that a significant frac-
tion of organic nitrogen is in a form of aliphatic
carbon, with methylamine (CH3NH2) and ethylamine
(CH3CH2NH2) being the two major carriers identified,
with similar abundances. These species are very volatile
and Sandford et al.
(2006) argue that they are likely
present in the cometary material in a form of an amine-
rich organic polymer rather than as a free primary
amine. One may speculate that energetic processing of
ices containing ammonia, methane, and ethane could
produce these molecules, as well as perhaps some or-
ganic polymers. Laboratory studies are urgently needed
to test this idea.
If such polymers are also a source
of cometary HNC, the low HNC abundance in comet
73P/Schwassmann-Wachmann appears consistent with
the observed depletion of ammonia and acetylene im-
plied by IR observations (Dello Russo et al. 2007). We
note that the rotational spectrum of methylamine in the
49 -- 326 GHz frequency range has now been measured in
the laboratory (Ilyushin et al. 2005). It is thus feasible
to search for this species in relatively active comets.
A comparison of the HNC/HCN ratio in comet
73P/Schwassmann-Wachmann with measurements in
other members of the 21P/Giacobini-Zinner composi-
tional class is of great interest. This class includes ob-
jects with extreme C2 and NH2 depletions that are re-
lated to C2H2 and NH3 depletions in nuclear ices. We
would expect a low HNC/HCN ratio in these objects as
well. So far, very few members of this class have been
identified (Fink & Hicks 1996). The upper limit for the
HNC/HCN ratio in Giacobini-Zinner obtained during its
1998 apparition (< 11%, Biver et al. 1999; Fig. 3) is not
sensitive enough to draw definitive conclusions. Comet
21P/Giacobini-Zinner will make a favorable apparition
in 2018 (at 0.4 AU from the Earth), during which sen-
6
Lis et al.
sitive measurements will be possible using the Atacama
Large Millimeter/Submillimeter Array (ALMA).
5. SUMMARY
Observations of hydrogen isocyanide in cometary at-
mospheres carried out to date indicate that the HNC
production has to be efficient in the inner coma, just
as the material leaves the nucleus. The process has to
be temperature dependent to explain the observed vari-
ation in the HNC/HCN ratio with the heliocentric dis-
tance. Thermal degradation of macromolecules or poly-
mers produced from ammonia and carbon compounds,
such as acetylene, methane, or ethane appears consistent
with all existing data, including the very low HNC/HCN
ratio in comet 73P/Schwassmann-Wachmann reported
here. Such polymers have been invoked previously to ex-
plain anomalous 14N/15N ratios measured in cometary
CN.
Additional
interferometric observations of HNC in
comets are needed to provide constraints on the spa-
tial distribution of this molecule in the cometary comae.
Such observations will soon be possible in moderately
productive comets with (sub)millimeter interferometers
(e.g., e-SMA). Measurements of the H15NC/H14NC iso-
topic ratio would be instrumental in determining whether
the HCN polymers similar to those invoked to explain the
enhanced 15N abundance in CN may also be a source of
cometary HNC. However, such measurements will have
to await the commissioning of ALMA.
This research has been supported by NSF grant
AST-0540882 to the Caltech Submillimeter Observatory.
D.C.L. acknowledges support from the Observatoire de
Paris during his stay in Meudon. S.B.C. acknowledges
support from NASA's Planetary Atmospheres Program.
We thank N. Fray for useful discussions and the referee,
M. DiSanti, for a thorough and helpful review of the
manuscript.
A'Hearn, M.F., Millis, R.L., Schleicher, D.G., Osip, D.J., & Birch,
Dello Russo, N., Vervarck, R.J., Jr., Weaver, H.A., Biver, N., et al.
P.V. 1995, Icarus 118, 223
2007, Nature, in press
Biver, N., Bockel´ee-Morvan, D., Boissier, J., Colom, P., et al. 2005,
in Asteroids, Comets, and Meteors, IAU Symp. No. 229 (Book
of Abstracts), 43
Biver, N., Bockel´ee-Morvan, D., Colom, P., Crovisier, J., et al. 1997,
Fink, U., & Hicks, M.D. 1996, ApJ 459, 729
Fray, N. 2004, Ph. Thesis, University of Paris XII
Herbst, E., Terzieva, R., & Talbi, D. 2000, MNRAS 311, 869
Hirota, T., Yamamoto, S., Mikami, H., & Ohishi, M. 1998, ApJ
REFERENCES
Science 275, 1915
Biver, N., Bockel´ee-Morvan, D., Crovisier, J., Colom, P., Henry, F.,
Moreno, R., Paubert, G., Despois, D., & Lis, D.C. 2002, EMP
90, 323
Biver, N., Bockel´ee-Morvan, D., Crovisier, J., Henry, F., et al. 2000,
AJ 120, 1554
Biver, N., Bockel´ee-Morvan, D., Boissier, J., Colom, P., et al. 2006a,
BAAS 38, 484
Biver, N., Bockel´ee-Morvan, D., Crovisier, J., Lis, D.C., et al.
2006b, A&A 449, 1255
Biver, N., Crovisier, J., Davis, J.K., Despois, D., et al. 1999, in
Asteroids, Comets, and Meteors 1999, p. 50 (Book of Abstracts)
Blackman, G.L., Brown, R.D., Godfrey, P.D., & Gunn, H.I. 1976,
Nature 261, 395
Bockel´ee-Morvan, D., Gunnarsson, M., Biver, N., Charnley, S.,
Crovisier, J., Lis, D.C., Rodgers, S. 2005, in Asteroids, Comets,
and Meteors, IAU Symp. No. 229 (Book of Abstracts), 111
Bockel´ee-Morvan, D. 2007, A&SS, in press (ALMA Symposium
Madrid, Spain)
Bockel´ee-Morvan, D., Biver, N., Moreno, R., Colom, P., et al. 2001,
Science 292, 1339
Boehnhardt, H., Holdstock, S., Hainaut, O., et al. 2002, EP 90, 131
Charnley, S.B. 2002,
in Proceedings of the Second European
Workshop on Exo/Astrobiology (ESA SP-518), 33
Charnley, S.B., & Rodgers, S.D. 2002, ApJ 569, L113
Charnley, S.B., Rodgers, S.D., Kuan, Y.-J., & Huang, H.-C. 2002,
ASR 30, 1419
Colom, P., Crovisier, J., Biver, N., Bockel´ee-Morvan, D., Boissier,
in SF2A-2006: Semaine de
J., & Lecacheux, A. 2006,
l'Astrophysique Fran¸caise, ed. D. Barret et al., 389
Crovisier, J., Biver, N., Bockel´ee-Morvan, D., Boissier, J., Colom,
P. et al. 2005, BAAS 37, 646
Crovisier, J., Biver, N., Bockel´ee-Morvan, D., Boissier, et al. 2006,
BAAS 38, 485
503, 717
Ilyushin, V.V., Alekseev, E.A., Dyubko, S.F., Motiyenko, R.A., &
Hougen, J.T. 2005, J. Mol. Spec. 229, 170
International Comet Quarterly 2006, Vol. 139, pp. 100, 116; Vol.
140, pp. 137,151, 174
Irvine, W.M., Bockel´ee-Morvan, D., Lis, D.C., Matthews, H.E.,
Biver, N., et al. 1996, Nature 383, 418
Irvine, W.M., Bergin, E.A., Dickens, J.E., Jewitt, D., et al. 1998,
Nature 393, 547
Irvine, W.M., Bergman, P., Lowe, T.B., Matthews, H., et al. 2003,
Origin of Life and Evolution of Biosphere 33, 609
Jehin, E., Manfroid, J. Cochran, A.L., Arpigny, C., et al. 2004, ApJ
613, L161
Lis, D.C., Keene, J., Young, K, Phillips, T.G., Bockel´ee-Morvan,
D., et al. 1997, Icarus 130, 355
Rodgers, S.D., & Charnley, S.B. 1998, ApJ 501, L227
Rodgers, S.D., & Charnley, S.B. 2001a, ApJ 546, 324
Rodgers, S.D., & Charnley, S.B. 2001b, MNRAS 323, 84
Sandford, S.A., Al´eon, J., Alexander, C. M. O'D., Araki, T., et al.
2006, Science 314, 720
Schilke, P., Walmsley, C.M., Pineau des Forets, Roueff, E., Flower,
D.R., & Guilloteau, S. 1992, A&A 256, 595
Snyder, L.E., & Buhl, D. 1972, Ann. New York Acad. Sci. 194, 17
Weaver, H.A., Lisse, C.M., Mutchler, M.J., et al. 2006, BAAS 38,
490
Villanueva, G.L., Bonev, B.P., Mumma, M.J., et al. 2006, ApJ 650,
L87
Zuckerman, B., Morris, M.P., Palmer, P., & Turner, B.E. 1972, ApJ
173, L125
|
astro-ph/9610258 | 1 | 9610 | 1996-10-31T00:04:25 | Cygnus X-1: A Case for a Magnetic Accretion Disk? | [
"astro-ph"
] | With the advent of RXTE, which is capable of broad spectral coverage and fast timing, as well as other instruments which are increasingly being used in multi-wavelength campaigns (via both space-based and ground-based observations), we must demand more of our theoretical models. No current model mimics all facets of a system as complex as an x-ray binary. However, a modern theory should qualitatively reproduce -- or at the very least not fundamentally disagree with - all of Cygnus X-1's most basic average properties: energy spectrum (viewed within a broader framework of black hole candidate spectral behavior), power spectrum (PSD), and time delays and coherence between variability in different energy bands. Below we discuss each of these basic properties in turn, and we assess the health of one of the currently popular theories: Comptonization of photons from a cold disk. We find that the data pose substantial challenges for this theory, as well as all other currently discussed models. | astro-ph | astro-ph |
Cygnus X-1: A Case for a Magnetic Accretion Disk?
Michael A. Nowak1, B. A. Vaughan2, J. Dove3, and J. Wilms4
Abstract. With the advent of RXTE, which is capable of broad spec-
tral coverage and fast timing, as well as other instruments which are
increasingly being used in multi-wavelength campaigns (via both space-
based and ground-based observations), we must demand more of our
theoretical models. No current model mimics all facets of a system as
complex as an x-ray binary. However, a modern theory should qualita-
tively reproduce -- or at the very least not fundamentally disagree with
-- all of Cygnus X-1's most basic average properties: energy spectrum
(viewed within a broader framework of black hole candidate spectral be-
havior), power spectrum (PSD), and time delays and coherence between
variability in different energy bands. Below we discuss each of these basic
properties in turn, and we assess the health of one of the currently pop-
ular theories: Comptonization of photons from a cold disk. We find that
the data pose substantial challenges for this theory, as well as all other
currently discussed models.
1. Energy Spectra
Historically, energy spectra have been the primary focus of both observers and
theorists. The low state spectrum of Cyg X-1 (see Fig. 1) for the most part is
well-fit by a power-law with a photon index ∼ 1.7. Additional facts to keep in
mind are: the total luminosity of Cyg X-1 is likely < 0.1 LEdd, and that most
other black holes show a strong soft component and softer power-law tail above
> 0.1 LEdd (cf. Nowak 1995). The softer, high state of Cyg X-1 is of roughly
comparable luminosity to the low state (Zhang 1996). The soft components that
do exist in the low state are weak compared to the power-law.
The most popular models for this state are hot, optically thin models [with
advection dominated disks (ADD) being one example, cf. Narayan et al. 1996],
and Comptonization of seed photons from a cold disk (cf. Haardt & Maraschi
1993; Dove, Wilms, & Begelman 1996). ADD theories provide an explanation
for the hard to soft transition at ∼ 0.1 LEdd, although they have yet to self-
consistently calculate the (weak) observed features that Comptonization mod-
1JILA, Campus Box 440, Boulder, CO 80309-0440
2Caltech, Pasadena, CA 91125
3also Dept. of APAS, University of Colorado, Boulder, CO 80309
4IAA University of Tubingen, Waldheuser Str. 64, Tubingen, Germany
1
Figure 1.
Best-fit corona model to unfolded, non-simultaneous Cyg
X-1 spectrum. Data are from BBXRT, TTM, (simultaneous with)
HEXE, and OSSE (Dove, Wilms, & Begelman 1996).
els attribute to reflection. One theoretical motivation for the Comptonization
models is the recent work (cf. Balbus, et al. 1996; Hawley et al. 1996) demon-
strating the existence of a powerful magnetic field-driven shear instability that
is strongest in low-density regions. Comptonization models assume that most
of the disk energy is dissipated in a corona, with roughly half being reprocessed
by the disk into soft photons (cf. Haardt & Maraschi 1993).
Fitting the spectra of Cyg X-1 above 2 keV typcally requires τes ∼ 0.2
and kTe ∼ 150 keV. With roughly half of the luminosity being reprocessed by
the disk, and a large fraction escaping the optically thin corona, we expect a
large soft flux at ∼ 200 eV. This is not observed (see Fig. 1 and Dove, Wilms,
Begelman 1996, where more refined models will be presented). In addition to not
offering an explanation for the low -- high transition at ∼ 0.1 LEdd, this is a severe
problem for the corona models. The models do, however, correctly account for
the Fe-line and reflection features, with ∼ half of the soft 200 eV excess being
due to reflection (as opposed to intrinsic disk emission). ADD models, with a
similar geometry and a strong source of hard photons, may also overproduce
soft X-rays if they attempt to model the observed Fe-line with reflection off of
the optically thick, geometrically thin outer disk.
2. Phase/Time Lags Between Soft and Hard Variability
Cyg X-1 shows strong X-ray variability (rms fluctuations ∼ 40%), with timescales
ranging from milliseconds to tens of seconds. This variability is usually measured
by Fourier transforming the observed time series and then forming the power
spectral density (PSD), i.e. the squared transform. One can also form the cross
power density (CPD) by multiplying together the complex transforms from two
time series. The Fourier frequency-dependent phase of the CPD is known as
the phase delay, θd(f ), from which one can form the time delay, τd ≡ θd/2πf .
Cygnus X-1 shows a characteristic time lag of the hard photons behind the soft,
with τd ∼ 2 (f /0.01Hz)−2/3 s (cf. Miyamoto et al. 1992). This corresponds to
a nearly constant phase delay of ∼ 0.1 rad.
2
)
.
c
e
s
(
g
a
L
d
r
a
H
100
10-1
10-2
10-3
10-4
10-2
kT
BB
= 0.15 keV
10-1
100
101
102
Fourier Period (sec.)
)
.
c
e
s
(
g
a
L
d
r
a
H
100
10-1
10-2
10-3
10-4
10-2
kT
BB
= 0.60 keV
10-1
100
101
102
Fourier Period (sec.)
)
.
c
e
s
(
g
a
L
d
r
a
H
100
10-1
10-2
10-3
10-4
10-2
kT
BB
= 1.50 keV
10-1
100
101
102
Fourier Period (sec.)
Figure 2.
Effects of Compton coronae on intrinsic phase lags, as a
function of input blackbody temperature and mean scattering path
length (λes = 106, 107, and 108 cm). From Nowak & Vaughan (1996).
The only currently proposed ADD model explanation for this delay invokes
thermal disturbances propagating from the outer disk edge toward the center (cf.
Manmoto et al. 1996). This model requires, however, a very slow propagation
speed ∼ 10−4c in order to be in quantitative aggreement with observations (cf.
Vaughan & Nowak 1996). Corona models explain the delay as possibly arrising
from the magnetic flares that are allegedly energizing the corona (cf. Nowak
1994). If the flare temperature increases throughout its evolution, hard photons
naturally lag soft photons. However, diffusion of photon scattering times within
the corona will lead to a "smearing out" of this intrinsic time lag, depending
upon the spectrum of the seed photons, the corona temperature, optical depth,
and mean free scattering path (cf. Fig.2; Miller 1995; Nowak & Vaughan 1996).
Current observations rule out large scattering paths (∼ 108 cm) with very soft
(∼ 150 eV) inputs. Upcoming RXTE observations (obtaining phase lags in 6-8
energy channels out to ∼ 250 Hz) will provide more stringent constraints.
3. Coherence Between Soft and Hard Variability
The coherence function is essentially the normalized average amplitude of the
CPD (cf. Vaughan & Nowak 1996, Nowak & Vaughan 1996). It is a measure of
the fraction of the two time series that can be interpreted as time-independent
and linear transforms of one another. Coherence must be less than or equal to
unity, and be equal to unity only when there is a linear, time-indpendent transfer
function relating one channel to the other. Coherence is less than unity under
any of the following conditions: variability is due to thermal flares that are dom-
inated by emission from the Wien tale; there are multiple, independent sources
of variability (whether linear or not) with multiple, independent responses to the
sources (whether linear of not); there is any non-linear transfer function from
one channel to the other (cf. Vaughan & Nowak 1996).
If coronae are truly formed via multiple magnetic flares, then we expect a
strong loss of coherence, unless the the response to each of the flares is identical.
Nowak (1994) showed that a kinematic model, with multiple flaring regions,
reproduced both the PSD and phase lags seen in the very high state of GX339 --
4, as well as qualititatively agreed with the energetics of coronal formation. This
model, and all others like it, produces a strong loss of coherence (Fig. 3). This
is not seen in Cyg X-1 (Fig. 3; Vaughan & Nowak 1996), which shows near
3
Figure 3.
Left: Coherence of theoretical model with multiple flares
(based on Nowak 1994). Right: Ginga data of Cyg X-1 (cf. Vaughan
& Nowak 1996) The Cyg X-1 loss of coherence is likely a noise artifact.
unity coherence between all well-observed energies. This also disagrees with the
model of Manmoto et al. (1996). For that case the disturbance shocks upon
reflection off of the disk inner edge, which is an inherently non-linear process,
and thus leads to less than unity coherence for a stochastic source of variability.
As they currently stand, coronal models of Cyg X-1 have several problems.
They predict too much soft X-ray flux, and they have yet to offer a viable
explanation for the phase lags that also preserves unity coherence. ADD models
also do not offer an explanation for the observed coherence, and they have yet
to self-consistently calculate the observed "reflection" and line features.
Acknowledgments. We would like to acknowledge useful conversations
with M. Begelman, P. Michelson, R. Staubert, C. Thompson, M. van der Klis,
and all those who weren't allowed to ask questions after the talk, but were
good enough to find me at coffee-break, especially F. Haardt and L. Maaraschi.
M.A.N. gratefully acknowledges support from NASA grant NAG 5-3225.
References
Balbus, S., et al. 1996, this volume
Dove, J., Wilms, J., & Begelman, M.C. 1996, ApJ, submitted
Haardt, F., & Maraschi, L. 1993, ApJ, 413, 507
Hawley, J.C.., et al. 1996, this volume
Manmoto, T., et al. 1996, ApJ, 464, L135
Miller, M.C. 1995, ApJ, 441, 770
Miyamoto, S., et al. 1992 ApJ, 392, L21
Narayan, R., et al. 1996, this volume
Nowak, M.A. 1994, ApJ, 422, 688
Nowak, M.A. 1995, PASP, 718, 1207
Nowak, M.A., & Vaughan, B.A., 1996 MNRAS, 280, 227
Vaughan, B.A., & Nowak, M.A., 1996 ApJ, submitted
Zhang, S.N., et al. 1996, this volume
4
|
astro-ph/9807180 | 2 | 9807 | 1998-08-27T19:06:29 | Distribution of Spectral Characteristics and the Cosmological Evolution of GRBs | [
"astro-ph"
] | We investigate the cosmological evolution of GRBs, using the total gamma ray fluence as a measure of the burst strength. This involves an understanding of the distributions of the spectral parameters of GRBs as well as the total fluence distribution - both of which are subject to detector selection effects. We present new non-parametric statistical techniques to account for these effects, and use these methods to estimate the true distribution of the peak of the nu F_nu spectrum, E_p, from the raw distribution. The distributions are obtained from four channel data and therefore are rough estimates. Here, we emphasize the methods and present qualitative results. Given its spectral parameters, we then calculate the total fluence for each burst, and compute its cumulative and differential distributions. We use these distributions to estimate the cosmological rate evolution of GRBs, for three cosmological models. Our two main conclusions are the following: 1) Given our estimates of the spectral parameters, we find that there may exist a significant population of high E_p bursts that are not detected by BATSE, 2) We find a GRB co-moving rate density quite different from that of other extragalactic objects; in particular, it is different from the recently determined star formation rate. | astro-ph | astro-ph | Distribution of Spectral Characteristics and the Cosmological Evolution of
GRBs.
Nicole M. Lloyd and Vah´e Petrosian
Center for Space Sciences and Astrophysics, Stanford University,Stanford, CA 94305-4060
ABSTRACT
We investigate the cosmological evolution of GRBs, using the total gamma-ray
fluence as a measure of the burst strength. This involves an understanding of
the distributions of the spectral parameters of GRBs as well as the total fluence
distribution - both of which are subject to detector selection effects. We present
new non-parametric statistical techniques to account for these effects, and use these
methods to estimate the true distribution of the peak of the νFν spectrum - Ep -
from the raw distribution. The distributions are obtained from four channel data
and therefore are rough estimates; hence, we emphasize the methods and present
qualitative results. Given its spectral parameters, we then calculate the total fluence
for each burst, and compute its cumulative and differential distributions. We use
these distributions to estimate the cosmological rate evolution of GRBs, for three
cosmological models. Our two main conclusions are the following: 1) Given our
estimates of the spectral parameters, we find that there may exist a significant number
of high Ep bursts that are not detected by BATSE, 2) We find a GRB co-moving
rate density quite different from that of other extragalactic objects; in particular, it is
different from the recently determined star formation rate.
Subject headings: gamma rays: bursts -- methods: statistical -- cosmology:
observations -- miscellaneous.
1.
Introduction
8
9
9
1
g
u
A
7
2
2
v
0
8
1
7
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Identification of several gamma-ray bursts (hereafter GRBs) with X-ray, optical and radio
afterglows, as well as measurements of the redshifts of GRB 970508, GRB 981214, and GRB
980703 are beginning to define the GRB paradigm. However, this information is not sufficient for
detailed cosmological studies of their distribution and evolution. Until more information on the
distances to GRBs are revealed, we must continue to rely on standard statistical studies (such as
obtaining number count distributions of fluence or flux) to gain further insight on the nature of
these events.
However, information on the spatial distribution - although necessary - will not be sufficient
for solving the puzzle of GRBs. In particular, we need spectral information to understand the
-- 2 --
energy release and radiation processes at the burst. This paper deals with the determination of
the spatial and spectral properties of GRBs.
The information on the spatial distribution comes primarily from the investigation of source
LK(z), where dL(z) is the luminosity distance, K(z) is the so-called K-correction
counts or the logN -logS relation. Assuming a luminosity function (LF) and a cosmological
model, one can convert such counts to spatial (or redshift) distributions (Weinberg, 1972).
The likelihood of the assumed luminosity function determines our confidence in the redshift
distribution or cosmological evolution of these sources. The usual practice is to use the peak flux
fp(ν) as a measure of the burst strength. This flux is related to the peak luminosity Lp(ν) via
Lp(ν) = fp(ν)Ωd2
due to redshift of the spectrum, and Ω is the solid angle into which the radiation is beamed
(Ω = 4π for isotropic emission). However, we have no knowledge of the burst peak luminosity
function - say Ψ(Lp(ν)) - and it is unclear what assumptions can be made about this function. For
example, in a fireball model, both the duration and the peak luminosities are expected to depend
on the details of the fireball (e.g. the bulk Lorentz factor and amount of baryon loading), so that
their distributions will be broad and perhaps comparable to the observed range of the peak fluxes.
As a result, the luminosity function cannot be readily de-convolved from the spatial distribution,
and we cannot get an accurate measure of the burst evolution or spatial distribution from the
logN -logS diagram.
On the other hand, in a neutron star (or black hole) merger model, the total energy released
- represented by the gravitational potential energy of the merging process - is expected to be well
defined and have a narrow distribution; this is also most likely true for the "hypernova" model.
Hence, we would like to determine what radiative signature is a representative measure of this
energy and the strength of the GRB.
As the name "gamma ray burst" implies, the peak gamma-ray flux fp (or luminsity Lp) of a
GRB far exceeds the peak fluxes at other energies. However, until recently, we did not know at
what frequency the total radiated energy peaked. The recent X-ray, optical and radio observations
of the afterglows - in particular the fact that these fluxes decline more rapidly than 1/t (see e.g.
Murakami et al., 1997, Galama et al., 1997) indicate that the energy fluence in the gamma-ray
range is also higher than in any other band. For isotropic emission or a beaming angle independent
of wavelength, this implies that the total gamma-ray radiated energy, E tot, is the best measure
of the total released energy. Therefore, it is plausible that the distribution of the gamma-ray
radiated energy is narrow, so that the total gamma-ray fluence Ftot = E tot(1 + z)/(Ωd2
L) provides
the best indication of the strength of the burst for the logN -logS analysis and other investigations
(assuming Ω has a narrow distribution). In particular, it is a better measure than the observed
fluence Fobs, which represents the fluence in some narrow energy range (i.e. the energy range in
which the detector triggers).
Note that the relation between the monochromatic flux (or fluence) and the corresponding
luminosities require a knowledge of the so called K-correction due to redshift of the spectrum.
-- 3 --
This complication is not present in the relation between the total radiated energy and the total
gamma ray fluence. It is also important to note that Ftot depends on the spectral properties of
the burst. Hence, if we want to simplify matters by dealing with the total fluence, we need to
understand the spectral distribution F (ν) of the GRBs.
The primary goal of this paper is to demonstrate how we can obtain information on the
cosmological evolution of GRBs, using the observed distribution of Ftot. However, for an accurate
determination of the distribution of Ftot, we need to determine the spectral parameters of each
burst. It is common to use the following four parameters (say, in a Band spectrum (1993)) to
characterize a burst's spectrum: the photon energy at the peak of the νFν spectrum, Ep, the low
energy spectral index, α, the high energy spectral index, β, and a normalization A. The best
source of data for the purpose of determining these parameters and Ftot is the BATSE instrument
on board the Compton Gamma-Ray Observatory (CGRO). However, because the BATSE trigger
criterion is based on photon counts between 50 - 300 keV (for the 3B catalog), the selection biases
or thresholds on Ftot as well as the spectral parameters can be quite complicated. In §2 we show
how the trigger criterion used by BATSE can be used to obtain the selection bias on any measured
quantity - in particular, Ftot and Ep - and describe new methods to account for the resultant
complicated data truncation. In §3, we present our results on the determination of the spectral
parameters. The distribution of Ep is given in §4, and in §5 we derive the distribution of Ftot, give
a scenario for cosmological evolutions of the GRBs, and discuss its relation to the new observations
of afterglows and evolution of other cosmological sources. A brief summary is presented in §6.
2. Description of Data and The Basic Problem
2.1. The Data
BATSE has produced a large catalog of GRBs with a wealth of information on their temporal
and spectral characteristics. This detector has been designed to optimally detect GRBs and
has been very successful in achieving this goal. Nevertheless, like all detectors, it has intrinsic
limitations and may miss certain types of bursts or provide only limited information on others.
BATSE is a triggered instrument; it will record data on a burst if the detector counts in a time
interval ∆t and energy range ∆E exceed some threshold value Cmin(∆t) determined by the
background fluctuations. BATSE uses three trigger time intervals ∆t = 1024, 256, and 64 ms with
four energy channels ∆E = 25 − 50, 50 − 100, 100 − 300, and > 300 keV. For most of the data,
triggering is based on counts in channels 2 and 3; recently, however, other channel combinations
have been tried. Here we will use the 4 channel LAD fluence values published in the BATSE
3B catalog (Meegan et al., 1995), which utilizes the former trigger scheme, so that the observed
50keV F (ν)dν. However, the formalism described below is applicable to the latter
data as well. Note that we analyze bursts with Cmax/Cmin on 1024ms timescale. Because of the
lack of resolution of the four channel data, here we will emphasize the method of our calculations
fluence Fobs ≡ R 300keV
-- 4 --
rather than give concrete quantitative results. Future work will involve calculations made from 16
channel CONT data.
2.2. Data Truncation
The BATSE detector trigger conditions lead to several limitations on the data. The most
obvious is that only bursts with peak counts Cmax > Cmin are detected. Determination of the
average V /Vmax = (Cmin/Cmax)3/2, and its distribution were one of the primary objectives of
BATSE using this trigger configuration (Meegan et al., 1992). Another limitation arises from the
finite durations ∆t of the triggering criteria. As pointed out by Petrosian and Lee (1994, see also
Lee and Petrosian, 1996), this reduces the detector efficiency of short duration bursts (duration
T < ∆t) in the sense that there is a relative bias against detection of weaker bursts with shorter
duration. Similarly, as pointed out by Piran and Narayan (1996), the finite range ∆E of the
detector energy channels provides another limitation (see also Cohen, Piran, and Narayan, 1998).
In this case, the detector efficiency is highest for bursts with most of their photons in the range
∆E, and decreases for bursts with more of their flux outside this range.
It turns out that we can account for these limitations by utilizing the BATSE trigger
condition: Let us consider a burst characteristic, say X, whose value is measured by BATSE. For
each burst, then, we know X, Cmax, and Cmin. Now, in the spirit of the V /Vmax test, one can ask
what is the possible range of values of X that this burst can have and still trigger the instrument.
In general, X can have an upper limit u, a lower limit l, or both limits; X ∈ T = [l, u]. This
question was first posed by Petrosian and Lee (1996) in connection with the determination of the
detection threshold or efficiency of the observed energy fluence Fobs; in this case, the fluence has
only a lower limit Fobs,lim. It was shown that this threshold (or for that matter, the threshold on
any other measure of burst strength which is proportional to the photon counts) is obtained from
the simple relation
Fobs/Fobs,lim = Cmax/Cmin,
(1)
provided the burst's spectrum does not change drastically throughout its duration. A similiar
relation holds for the total fluence Ftot as we describe in §5.
For other bursts characteristics, this may not be the case. For example, we have shown (Lloyd
and Petrosian 1998) that the spectral parameter Ep has both an upper and lower limit. The
values of Ep,max and Ep,min are not related to Cmax/Cmin and Ep in a simple form as above, and
thus we require a more complex procedure to determine their values. We will describe this in §3.
Given the thresholds, we can then use non-parametric statistical techniques to obtain an estimate
the parent distribution of the relevant observable from the observed distribution.
2.3. Dealing with Data Truncation
-- 5 --
Correcting for data truncation in the one-sided case was discussed in detail by Efron and
Petrosian (1992). This method has been used to determine the correlation between parameters
in a bivariate (or multivariate) setting, and to obtain univariate distributions of several GRB
characteristics (see, e.g., Lee and Petrosian, 1996). The basic idea is that one can use information
from the untruncated region of a distribution to estimate how much data is missed due to the
truncation. For a description of the basic process, see also Petrosian (1992 or 1993).
Dealing with data which has both upper and lower limits is more complex and requires a
generalization of the above technique. Recently, Efron and Petrosian (1998) have developed a
method to deal with this kind of data truncation. We briefly review this method here.
Consider data points (xi, yi) (in our case (Ep, Fobs)), where xi has lower and upper limits li
and ui, respectively; xi ∈ Ti = [li, ui]. Let us also assume that xi and yi are uncorrelated. Let
g(x) be the true distribution of x, which would be observed if there were no truncations. However,
because xi is limited to Ti, we observe the conditional distribution g(xiTi) = g(xi)/G(Ti), where
G(Ti) = [Pj g(xj ) : xj ∈ Ti] is the probabilitly that x exists in Ti. We normalize g(xi) such that
PN
i=1 gi = 1, and define the vectors
and the matrix
gi = g(xi) and Gi = G(Ti)
Jij = (cid:26) 1, xj ∈ Ti,
0, xj /∈ Ti.
(2)
(3)
Furthermore, by definition, one can show that Gi = Jijgj (where we have used the convention that
repeated indices are summed over). The goal is to estimate g(x) from li, ui, and xi assuming all N
cases are independently distributed. It can be shown that maximizing the likelihood (see Maloney
et al., 1998) of the observed data set gives
1/gi = Jji ∗ (1/Gj ) + const,
(4)
where Jji is the transpose of Jij. The procedure amounts to solving this equation iteratively, under
the normalization condition and definition of Gj and Jij given above. Convergence is reached
when const in the above equation goes to zero. This method can be used to determine univariate
cumulative (G) and differential (g) distributions. As described in Efron and Petrosian (1998), it
can also be used to determine correlations between relevant variables (e.g. fluence and Ep or peak
flux and Ep).
We will use the methods described above to correct for data truncations and produce bias
free distributions of the peak energy Ep and the total fluence Ftot; we will then directly use the
latter distribution to investigate the cosmological evolution of GRBs.
3. Spectral Characteristics
-- 6 --
The spectral properties of GRBs provide the most direct information about the physical
processes associated with the event. In particular, the distribution of the spectral parameters
and the correlation between these and other GRB characteristics can shed significant light on the
radiation mechanisms and energy production of the burst.
Therefore, our first task is to determine the spectral parameters of each burst. We characterize
the spectral forms of the bursts by four parameters - α and β (the photon power law index at
low and high energies respectively), Eo (a break energy related to the peak of the νFν spectrum)
and A (a normalization in units of ergs/cm2). We choose two functional forms to represent the
energy fluence spectra - a Band spectrum, and a generic double power-law spectrum with a
smooth transition between the low energy and high energy power law behavior:
F (E) = ( (A/Eo)(E/Eo)α+1exp[(β − α)(E/Eo − 1)], E < Eo
(A/Eo)(E/Eo)β+1,
E > Eo
F (E) =
(A/Eo)(E/Eo))α+1
1 + (E/Eo)α−β
(5)
(6)
We are interested in bursts whose νFν spectra have a maximum, so that we may correctly speak
of Ep as the "peak of νFν ". To define Ep (as well as to keep the burst's total fluence Ftot finite),
we require α to be greater than -2 and β to be less than -2 (which also implies α > β). As we
discuss below, this does not impose any large bias in our sample. Then, with these constraints, we
can define Ep for each spectrum above: Ep = Eo( α+2
β−α for (5) and (6),
respectively.
α−β ) and Ep = Eo( −(β+2)
α+2 )
1
From four channel data, we can determine the four spectral parameters with no degrees of
freedom. However, this introduces strong interplay between Ep and β or α. We avoid this by
selecting β from an assumed gaussian distribution with a mean of ≈ −3, and a standard deviation
of 1 (and of course the constraint β < −2). This allows for one degree of freedom. In fact, we
have performed this analysis using several distributions for β (i.e. a uniform distribution, or
a gaussian with a different mean and standard deviation), and have found similiar qualitative
results. However, we should note that steeper, more negative β's will naturally produce higher
Ep's, to accommodate the fluence in high energy channels (3 and 4). We find the rest of the
parameters by minimizing χ2 via a downhill simplex routine (e.g. see Press et al., 1992). For
the Band spectrum, we find acceptable fits for 291 bursts, while 268 give acceptable fits for the
second spectrum listed above, out of a total of 518 bursts. Figures 1(a) and 1(b) show the raw
distributions of the spectral parameters obtained with this method for the two spectral forms in
equation 5 and 6, respectively.
It turns out that the distribution of the parameters for bursts with unacceptably high χ2 is
not very different from those with acceptable χ2. We have tested to see if the reason for many high
χ2 bursts came from the constraints on α and β mentioned above. Relaxing these restrictions, we
found that only ∼ 1% of bursts that gave acceptable fits had an α < −2, β > −2, or α > β. We
-- 7 --
therefore assume that the bursts with good fits are representative of the total sample. However,
because the lack of energy resolution of the four channel data, we would like to emphasize that
our fits are rough estimates of the actual spectral parameters, so that the following analysis
exhibits our method rather than presenting a final quantitative result. As mentioned previously,
in future work we will use fits from 16 channel CONT data, which will provide a more accurate
determination of these parameters.
Several studies have obtained the spectral parameters for the brightest GRBs (e.g., Band et al.
1993) and investigated correlations between these parameters and peak flux, duration and spatial
distributions (Mallozzi et al. 1995, Kouvelioutou et al. 1995, Pendleton et al. 1996). However,
caution is required in such studies, when dealing with the raw data. The effects of selection biases
must be evaluated before obtaining these distributions and correlations. The parent distributions
of the spectral parameters could be quite different from the observed distributions shown in
the above figures. The differences are caused by the truncations resulting from the triggering
procedure, and are a measure of trigger efficiencies for each of the four spectral parameters
α, β, Ep, and A or Ftot = Aξ(α, β). The truncation on A or Ftot is described by equation (1):
Ftot,lim = Ftot(Cmin/Cmax). As mentioned previously and shown in detail in the next section,
there are also selection biases against high and low values of Ep. For instance, BATSE is most
likely to see bursts with Ep in the energy range in which the detector triggers. Becuase the
detector needs some minimum number of counts Clim to trigger, then for each burst one can find
an interval [Epmin, Epmax] to which Ep is confined in order for the burst to have been detected.
Similarly, truncation is present for α and β. For example, if a burst with given Ep and A had
a value of α which was postitive and large or had a value of β which was negative and large, -
i.e.
if the νFν spectrum had a steep rise or fall - then the observed Cmax could fall below the
threshold Cmin and render the burst undetectable. Using the methods described in §2 and below,
we can correct the raw distribution of all four parameters and determine their correlations with
other GRB characteristics. However, the effect of the truncation on α and β is more complicated
and not as pronounced as that of Ep and Ftot. We will deal with the former in future publications.
Here, we limit our discussion to obtaining the true distributions of Ep and Ftot. We will present
the results from the Band spectrum fits to the bursts only; the smooth double power-law spectrum
in equation (6) gives qualitatively similiar results for all of the following analysis.
4. Distribution of Ep
For each burst we have the values of α, β, A, Ep, and the observed fluence
Fobs = R E2
E1 Fα,β,A(E, Ep)dE, with E1 = 50keV and E2 = 300keV . From equation 1, we
can calculate the fluence threshold Fobs,lim. Using this limit, we can determine the possible range
of values Ep can take on, and still trigger the BATSE instrument. It is clear that the limiting
-- 8 --
value Ep,lim must satisfy the equation
Fobs,lim = Z E2
E1
Fα,β,A(E, Ep,lim)dE.
(7)
For the spectral forms and parameter ranges discussed in the previous section, the right hand
side of equation (7) increases monotonically with Ep,lim, reaches a maximum for Ep,lim within
the limits of the integration, and then decreases monotonically. Consequently, there will be two
values of Ep,lim that satisfy this equation. To find these two solutions, we start with the value
of Ep determined by the simplex routine (for which the right side is equal to Fobs), and increase
it (while keeping α, β, and A equal to their determined values) until equation (7) is satisfied -
that is, until the observed fluence is brought below the threshold. This value is the upper limit
to Ep, Ep,max. We then decrease Ep until this equation is again satisfied. This gives the lower
limit Ep,min. In essence, the procedure amounts to redshifting/blueshifting the burst until it is
undetectable by BATSE. Figure 2 shows Ep vs. Ep,max and Ep,min for the Band spectrum. Notice
the truncation is much more prominent on the upper than the lower end; hence, we expect to see
a more significant correction to the high end of the Ep distribution.
Using the method described in §2.3 for doubly truncated data, we corrected the observed
distribution given that the observations can detect only bursts with Ep limited to the interval
[Epmin, Epmax]. Figure 3 shows the raw and corrected differential distribution of Ep for the
Band spectrum. The raw and corrected distributions are normalized such that they are equal
at low values of Ep (where the data truncation is inconsequential). This normalization visually
emphasizes the differences of the distributions at high Ep. The figures suggest that a large number
of high Ep bursts are missed by the triggering procedure.
In order to determine the significance of the difference between the raw and corrected
distributions, we carry out the following two checks: First, we perform a K-S test on the
cumulative distributions of Ep, and a χ2 test on their differential distributions for both the Band
spectrum and the smooth double power law. For the χ2 test, we divide the distributions into two
bins, with equal numbers of observed bursts per bin. As shown in the table below, in both cases
the tests indicate the probability that the observed and corrected Ep distributions are the same is
extremely small.
Spectrum
Test
K-S
χ2
K-S Double Power Law
χ2
Double Power Law
Band
Band
Probability the distributions same
1.8 × 10−6
3.0 × 10−7
4.4 × 10−5
2.4 × 10−5
Second, we have carried out simulations to determine the accuracy of the method described in
§2.3; Appendix A contains these results. We begin with a parent distribution of Ep's, simulate an
-- 9 --
observational truncation (defined by the limiting fluence of the burst; see §4). From this we obtain
our observed distribution of Ep's. We then apply the above technique to correct for the truncation,
and retrieve our parent distribution to very high accuracy, demonstrating the robustness of the
method (see Appendix A for more details).
Our results are in qualitative agreement with the observational data from the Solar Maximum
Mission (SMM) presented by Harris and Share (1998a, 1998b); SMM is sensitive to higher energies
than BATSE, so that it suffers the bias against detection of bursts with hard spectra to a much
lesser degree. The results presented in Harris and Share (1998b) agree qualitatively with our
corrected Ep distribution. We perform a similiar χ2 as above, comparing the Ep distribution from
SMM with both the uncorrected and corrected BATSE Ep distributions presented in Figure 3.
In this case, we divide the data into two bins below and above 550 keV. We find the probability
that the Ep distribution from SMM data is the same as the raw BATSE Ep distribution to be
1.3 × 10−6. On the other hand, we find a 90% probability that the Ep distribution from the SMM
data is the same as the corrected BATSE Ep distribution. Note that the statistical method we
use can correct the Ep distribution only for the values of Ep observed by BATSE. Therefore, we
cannot determine the distribution above ∼ 1 MeV (where the raw BATSE distribution ends).
Similarly, because we have no observed Ep's below ∼ 20 keV, we cannot confirm Strohmeyer et al.
(1998) reporting a significant number of low Ep bursts (< 20 keV) observed by GINGA, but not
observed by BATSE.
A bias against detection of high Ep bursts has significant implications on correlations of Ep
with other measures of burst strength like peak flux, Fobs, or Ftot. The detector is most likely to
miss the high Ep bursts with low strength, so that a simple correlation analysis between raw values
of Ep and burst strength without the consideration of the data truncation can lead to erroneous
correlations. Examples of studies which may be affected by such a bias include the correlation
studies of Nemiroff et al. (1994) and Mallozzi et al. (1995). This will also have an important
effect on determining whether or not the correlations are intrinsic or due to cosmological redshift
(Brainerd, 1997). The above results indicate that a more thorough analysis of the correlations is
required before conclusions can be reached on the redshifts or intrinsic effects. We will address this
question in a future publication, when we have more accurate estimates of the spectral parameters.
5. Distribution of the Total Gamma Ray Fluence
As discussed in the introduction, an important way to learn about the spatial distribution
of GRBs is to study the distribution of some standard candle variable of the event. We also
conjecture the total radiated energy E tot in the gamma-ray range may be the best candidate for
such a standard candle variable, and that the total fluence Ftot of the GRB may provide the best
measure of distance. Consequently, we study the distribution of the bursts' total fluence; if the
-- 10 --
burst radiates isotropically, the total fluence is related to the total radiated energy E tot as:
Ftot =
Etot
4πd2
L
(1 + z),
(8)
where dL, the luminosity distance, depends on redshift and the cosmological model (Weinberg,
1972).
Again, because the detector is sensitive only over a finite energy range, BATSE does not
necessarily obtain all photons from the burst. Hence, the detector measures only a portion
of the burst's total fluence. However, the total fluence of a GRB can be obtained from the
spectral fits to the observed fluence simply by integrating over the spectrum of that burst:
Emin
Ftot = R ∞
Fα,β,A(E, Ep)dE., where Emin denotes the beginning of the gamma ray energy range
(as an estimate, we use Emin = 25 keV). This fluence will have an observational lower limit in the
same way that the fluence from 50 − 300 keV has, and its threshold is given by the generalization
of equation (1): Ftot,lim = Cmin
Ftot. Following the steps described in Petrosian and Lee (1996), we
Cmax
first test for any correlation between Ftot and Ftot,lim and parametrically remove this correlation.
We then use the method described in Efron and Petrosian (1994) for one sided truncated data (this
is equivalent to the method of §2.2, taking the upper limit u to ∞), and obtain the cumulative
and differential distributions for the total fluence.
Figures 4(a) and 4(b) show the cumulative and differential distributions of the total and
observed fluence for our sample. Marked on each curve are GRB 970508 (circle), GRB 971214
(cross), and GRB 980703 (triangle). Note that we do not find a good spectral fit for GRB 980703;
hence, we obtain a lower limit to its total fluence simply by summing up the published fluence
values for the four channels of the LAD detector. Following the tradition of radio astronomers (see
Lee and Petrosian, 1996), we divide out the −3/2 power law dependence of the number on the total
fluence. In this way, deviation from the horizontal line suggests deviation from a homogeneous,
isotropic, static, Euclidean distribution. In these figures, we have included all 518 bursts available
- including those whose spectral fits gave high values of χ2 - so as not to underestimate any part
of these distributions. We have also applied the same procedure to the subsample of bursts with
acceptable values of χ2 and find that the distributions are similiar to those of the complete sample
within about 30 %. Note that the fluence distributions in Figures 4(a) and 4(b) are unlike the
counts of other well known extragalactic sources such as galaxies, radio sources and AGNs or
quasars. The transition from 3/2 power law is too abrupt and the slope beyond this transition is
nearly constant. Clearly, some extraordinary evolutionary processes are at work. We explore this
in more detail in the next section.
5.1. Rate Evolution
There have been several detailed analyses of the logN -logS distribution for the peak fluxes of
GRBs with inconclusive results. This is primarily because any observed distribution can be fitted
-- 11 --
to an arbitrary luminosity function and evolution even if one assumes a cosmological model (see
e.g. Rutledge et al. 1995, Reichert and M´esz´aros, 1997, M´esz´aros and M´esz´aros, 1995, Hakkila et
al., 1996). Additional uncertainty associated with these results comes from neglect of the time
bias, spectral bias, and uncertainty in the value of the K-correction. These source of error are
absent when dealing with the total fluence, which requires no correction for time bias (Lee and
Petrosian, 1996), includes the spectral bias (described above), and requires no K-correction. Thus,
the relation between the total fluence counts and either the cosmological models or the luminosity
function and its evolution is more straight forward. Here we concentrate on the rate evolution of
GRBs.
To obtain some indication of possible evolution, we assume several representative values for
the total radiated energy E tot and derive the comoving rate density from the observed differential
counts of the fluences as
ρ(z) = (1 + z)n(Ftot)(dV /dz)−1(dFtot/dz)
(9)
where n(Ftot) is the differential distribution of the total fluence. Note that in all of our calculations,
we have assumed Ho = 60km s−1Mpc−1
Figures 5(a), 5(b) and 5(c) show ρ(z) for three representative cosmological models. In order
to span the range of possible evolutions, we choose three extreme models: a flat matter dominated
model (Ωm = 1, ΩΛ = 0), a curvature dominated model (Ωm = 0.2, ΩΛ = 0), and a flat vacuum
energy dominated model (Ωm = 0, ΩΛ = 1). Here, Ωm is the density parameter (equal to the
ratio of matter density to the critical density), and ΩΛ = Λc2/3H 2
o (where Λ is the cosmological
constant). Within each figure, ρ(z) is plotted for four standard candle energies: E tot = 1051, 1052,
1053, 1054 ergs. Superposed on each figure is the star formation rate (SFR) from Madau et. al
(1997, solid line), this rate delayed by 2 × 109 years (dashed-dot line), as well as the SFR convolved
with a distribution of time delays P (t) ∝ t−1 (dotted curve), which may arise in a neutron star or
black hole merger scenario (see e.g. Totani, 1997). Note that the SFR was calculated using an
Ωm = 1, ΩΛ = 0 cosmology (Madau, 1997); however, the shape of this curve as well as the 2 other
curves derived from the SFR are fairly insensitive to the cosmological model (see e.g., Totani,
1997).
Also marked on each figure are the rate densities of GRB 970508 (circle), GRB 971214
(cross) and GRB 980703 (triangle), given their measured redshifts, and the energy E tot required to
produce a burst at this redshift (given its total fluence). Table II lists this required energy. Again,
because we have only obtained a lower limit to the total fluence of GRB 980703, we only list a
lower limit to its required radiated energy.
-- 12 --
Burst
GRB 970508
GRB 970508
GRB 970508
GRB 971214
GRB 971214
GRB 971214
GRB 980703
GRB 971214
GRB 971214
Redshift Ftot ergs/cm2
Cosmology
Required E tot ergs
0.83
0.83
0.83
3.43
3.43
3.43
0.97
0.97
0.97
7.0 × 10−6
Ωm = 1 ΩΛ = 0
7.0 × 10−6
Ωm = 0.2 ΩΛ = 0
7.0 × 10−6
Ωm = 0 ΩΛ = 1
1.2 × 10−5
Ωm = 1 ΩΛ = 0
1.2 × 10−5
Ωm = 0.2 ΩΛ = 0
1.2 × 10−5
Ωm = 0 ΩΛ = 1
> 6.1 × 10−5
Ωm = 1 ΩΛ = 0
> 6.1 × 10−5 Ωm = 0.2 ΩΛ = 0
> 6.1 × 10−5
Ωm = 0 ΩΛ = 1
1.0 × 1052
1.4 × 1052
2.5 × 1052
1.7 × 1053
4.4 × 1053
1.8 × 1054
> 1.1 × 1053
> 1, 6 × 1053
> 3.2 × 1054
In all cases, the curve corresponding to E tot = 1051 ergs is ruled out by the observational
data. The E tot ≥ 1052 ergs curve accommodates GRB 970508, while GRB 971214 and GRB
980703 require E tot > 1053 ergs. In the ΩΛ = 1 case, the required E tot > 1054 ergs for GRB 971214
begins to exceed theoretical limits, which may be evidence against this cosmological model or
evidence for strong beaming of the burst radiation. Our results suggest that the total gamma-ray
radiated energy is not a standard candle and is spread over at least one decade. We emphasize
that these results depend critically on the total fluence, which in turn depends on accurate values
for the spectral parameters for these bursts. Higher resolution fits to the data as well as more
measurements of the redshifts of bursts will allow us to constrain the curves more definitely.
Finally, note that none of the three curves for ρ(z) follow the star formation rate curve, as
one would expect in many GRB models such as the merger or hypernovae models. The density
evolution for GRBS in unlike that for AGN's or ordinary galaxies as well. The results agree - at
least qualitatively - with those of Totani (1998). However, this is in contradiction with recent
claims that the GRB rate follows the star formation rate (Wijers et al., 1998).
6. Summary and Conclusion
The primary aim of this paper has been to investigate the cosmological evolution of GRBs,
which involves understanding the distributions of the spectral parameters as well as the total
fluence of the burst. Selection effects limit the information we can observe to explore these
distributions.
We have presented new non-parametric methods to account for these effects. We have applied
these methods first to the distribution of the peak of the νFν spectrum, Ep. We find that this
spectral characteristic suffers both upper and lower thresholds, although the upper threshold has
the dominant effect. We correct the observed distribution for this truncation and present the
parent distribution of Ep. This contains a large number of high Ep bursts not evident in the
observed distribution, which is in qualitative agreement with the SMM results of Harris and Share
-- 13 --
(1998b). This is important for studying correlations between Ep and other burst characteristics,
as well as for theoretical considerations concerning GRB models.
Using the spectral fits for each burst, we determine a rough measure of the total gamma-ray
fluence and present the GRB source counts based on this measure. We obtain both the differential
and cumulative counts of the total and observed fluences.
We convert these distributions to the comoving rate density of GRBs for 3 different
cosmological models and 4 different values of the total gamma ray radiated energies. We find
that none of the curves follow the star formation rate as predicted by various GRB models. More
observations on the redshifts of GRBs as well as high resolution spectral data (to obtain a precise
value for the total fluence) are needed to substantiate these results.
As a final note, we summarize some of the caveats in our analysis. First, the spectral
parameters are determined based on four channel data and are only rough estimates of the true
spectral properties of the burst. In particular, because β is chosen from an assumed distribution,
the correction to the Ep distribution could be overemphasized. These spectral parameters also
affect the values of the total fluence (since Ftot = R Fα,β,A(E, Ep)dE), which in turn affect the
density evolution functions for the bursts. Furthermore, to derive the density curves, we have
assumed that the bursts total fluence is a standard candle; although this seems to be the most
plausible candidate for a standard candle, the observations of GRB 970508, 971214, and GRB
980703 indicate the contrary. Note, however, that we can use a standard candle upper and lower
limit to at least constrain the possible evolution of GRBs. Finally, effects such as beaming were
not taken into account in the density evolution analysis. For a burst beamed into a solid angle Ω,
the total radiated energy E tot will decrease by a factor of Ω/4π from the isotropic total radiated
energy; hence, the observed dispersion in the total radiated energy (about an order of magnitude
or higher) could be due to variation in the degree of beaming. Meanwhile, beaming causes the
number of burst occurances over some time interval to increase by a factor of 4π/Ω. We will
address the issue of how beaming can affect the cosmological rate evolution in a future publication.
We would like to acknowledge P. Madau for the star formation rate data used to generate
the three SFR curves in 5(a),5(b), and 5(c). We would also like to thank B. Efron for help in
implementing the statistical methods discussed in the text. Finally, N.M. Lloyd would like to
thank members of the BATSE team for many useful discussions, and their hospitality during her
visit to MSFC.
A. Appendix
Here, we describe the results of simulations which exhibit the technique and demonstrate the
accuracy of the procudure used for correcting doubly truncated data. For this analysis, we use
the Band spectrum displayed in equation 5. We pull Ep from a uniform distribution between 5
and 35 keV. α is also drawn from a uniform distribution ranging between -0.9 and 5.9, as was β
-- 14 --
which ranged from -9.0 to -2.1. Finally, our normalization parameter A is taken from a uniform
distribution from 12 × 10−9 to 18 × 10−9ergs/cm2. Note that we are not trying to simulate the
real distribution of any of these parameters - we choose these distributions for the purpose of
displaying selection effects and to what degree our method can account for them.
Once we have the spectral parameters, we can calculate the fluence in each of the four BATSE
LAD energy channels. The limiting fluence of each burst is chosen from a gaussian distribution
with a mean of 6 × 10−11ergs/cm2 and a standard deviation of 6 × 10−10ergs/cm2 (note again, this
is not an attempt to simulate the actual BATSE detector response). We then ask the question:
How many of the 1000 simulated bursts have their fluence in channels 2 and 3, greater than the
limiting fluence? We find 554 bursts survived this cut - we will call this the observed sample.
We then proceed to apply the method described in §2 to our observed sample, to see if we could
get back the original distribution of Ep's. Figure 6 shows the the parent sample, the observed
sample, and the correction to the observed sample. As evident, the correction accounts for the
truncated data, and approximately reproduces the original sample. See §2 for a full description of
the technique.
-- 15 --
REFERENCES
Band, D., et al. 1993, ApJ, 413, 281
Brainerd, J.J. 1997, ApJ, 487, 96
Cohen, E., Piran, T., Narayan, R. 1998, ApJ, 500, 888
Efron, B. & Petrosian, V. 1992, Ap.J., 399, 345
Efron, B. & Petrosian, V., JASA, in press
Galama, T., et al. 1997, in Gamma Ray Bursts, AIP Conf. Proc. 428, eds. C.A. Meegan, R.D.
Preece, T.M. Koshut (New York: AIP) , 478
Hakkila, J., et al. 1996, in Gamma Ray Bursts, AIP Conf. Proc. 307,eds. G.J. Fishman, J.J.
Brainerd, K.Hurley (New York: AIP), 387
Harris, M.J. & Share, G.H. 1998a, in Gamma Ray Bursts, AIP Conf. Proc. 428, eds. C.A. Meegan,
R.D. Preece, T.M. Koshut (New York: AIP), 314
Harris, M.J. & Share, G.H. 1998b, ApJ, 494, 724
Horack, J.M., Emslie, A.G., Hartmann, D.H. 1995, ApJ, 447, 474
Kouveliotou, C., et al. 1995, in Gamma Ray Bursts, AIP Conf. Proc. 384, eds. C Kouveliotou,
M.S. Briggs, G.J. Fishman (New York: AIP), p.42
Kulkarni, S., et al. 1998, Nature 393, 35
Lee, T. & Petrosian, V. 1996, ApJ, 470, 479
Lloyd, N.M & Petrosian, V. 1997, in Gamma Ray Bursts, AIP Conf. Proc. 428, eds. C.A. Meegan,
R.D. Preece, T.M. Koshut (New York: AIP) , 67
Madau, P. 1997, astro-ph/9709147
Mallozzi, R.S., et al. 1995, ApJ, 454, 597
Maloney, A. & Petrosian, V. 1998, ApJ, in press
Meegan, C., et al. 1992, Nature, 355, 143
Meegan, C., et al. 1995, ApJS, 106, 65
M´esz´aros, P. & M´esz´aros, A. 1995, ApJ, 449, 9
Metzger et al. 1997, Nature, 387, 879
-- 16 --
Murakami, T. et al. 1997, in Gamma Ray Bursts, AIP Conf. Proc. 428, eds. C.A. Meegan, R.D.
Preece, T.M. Koshut (New York: AIP) , 435
Nemiroff, R.J., et al. 1997, ApJ, 435, L113
Pendleton, G.N., et al. 1996, ApJ, 464, 606
Petrosian, V. 1992, in Statistical Challenges in Modern Astronomy, eds. E.D.Feigelson and G.J.
Babu (New York: Springer-Verlag), 173.
Petrosian, V. 1993, ApJ, 402, L33
Petrosian, V. & Lee, T. 1996, ApJ, 467, L29
Piran, T. & Narayan, R. 1995, in Gamma Ray Bursts, AIP Conf. Proc. 384, eds. C Kouveliotou,
M.S. Briggs, G.J. Fishman (New York: AIP), 233
Press, W. H., Teukolsky, S. A., Vetterling W. T., and Flannery,B. P. 1992, Numerical Recipes in
FORTRAN (Cambridge: Cambridge University Press)
Reichert, D.E. & M´esz´aros, P. 1997, ApJ, 483, 597
Rutledge, R.E., et al. 1995, MNRAS, 276, 753
Strohmeyer, T.E., et al. 1997, astro-ph/9712332
Totani, T. 1997, ApJ, 486, L71
Totani, T. 1998, astro-ph/9805263
Weinberg, S. 1972, Gravitation and Cosmology: Principles and Applications of the General Theory
of Relativity (New York: Wiley)
Wijers, R.A.M., et al. 1998, MNRAS, 294, L13
This preprint was prepared with the AAS LATEX macros v3.0.
-- 17 --
30
20
10
0
20
15
10
5
0
0.1
1
10
-6
-4
-2
20
15
10
5
0
30
20
10
0
1
10
30
20
10
0
0.1
25
20
15
10
5
0
15
10
5
0
50
40
30
20
10
0
-6
-4
-2
-2
-1
0
1
2
0.0001
-2
-1
0
1
2
0.0001
Fig. 1. -- The distribution of the spectral parameters determined from a) the Band spectrum and b)
a smooth double power-law spectrum respectively. Note that β is drawn from the same truncated
gaussian distribution in both cases, and is not a fitted parameter.
100
10
1
1
0.1
Band Spectrum
1
1
10
10
Fig. 2. -- The maximum and minimum values of Ep for the Band spectrum. Note that the
truncation is more severe from above than below, which indicates that observations may miss
a population of GRBs with high Ep.
-- 18 --
10
Band Spectrum
Corrected
Uncorrected
1
1
10
Fig. 3. -- The observed and "corrected" distribution of Ep for the Band spectrum. The figure
suggests that there is a significant sample of high Ep bursts undetected by BATSE. As mentioned
in the text, the methods we used will not produce a correction beyond the raw observed distribution.
Fig. 4. -- The cumulative (a) and differential (b) distribution of both the observed (Fobs) and total
(Ftot) fluences. The circle indicates the observed and total fluences of GRB 970508, the cross marks
GRB 971214, and the triangle marks GRB 980703. In each figure, note the similarities between
the two distributions; they both display an abrupt break from the HISE dependence (see text),
indicating presence of strong evolutionary processes.
-- 19 --
10
10
1
0.1
0.01
1
1
10
10
1
0.1
0.01
10
1
0.1
0.01
0.001
1
10
Fig. 5. -- The comoving rate density versus redshift z for 3 different cosmological models indicated
by the values of Ωm and ΩΛ. Within each plot, we show the rate density for 4 different standard
candle energies; the numbers 0.1, 1, 10, and 100 mark the total gamma-ray radiated energy in
units of 1052 ergs. Note that these curves are significantly different from: 1) the star formation rate
(SFR) from Madau et al., 1997 (solid line), 2) the star formation rate with a time delay of 2×109 yr
(dot-dash line), and 3) the star formation rate convolved with a distribution of time delays (dotted
line, Totani et al., 1997). GRB 970508 (circle), GRB 971214 (cross) and GRB 980703 (triangle)
are indicated, given their measured redshifts and the energy E tot required to produce a burst at
this redshift (given its total fluence, see text). Note that no single standard candle energy can
accomadate the observed redshifts of these bursts, in any of the three models.
-- 20 --
2.5
2
1.5
1
0.5
0
Original Distribution
Observed Distribution
Corrected Distribution
10
Fig. 6. -- The original Ep distribution (solid line), the observed Ep distribution given the criterion
Fobs > Flim (dashed line), and the corrected distribution using the method described in §2 of the text
(dotted line). These distributions are obtained from the simulations described in the Appendix.
Note that the procedure recovers the original distribution almost exactly. Any differences are
primarily due to differences in the binnings of the distributions.
|
astro-ph/0506490 | 2 | 0506 | 2005-07-11T09:59:19 | Rest-frame optical and far-infrared observations of extremely bright Lyman-break galaxy candidates at z ~ 2.5 | [
"astro-ph"
] | We have investigated the rest-frame optical and far-infrared properties of a sample of extremely bright candidate Lyman-break galaxies (LBG) identified in the Sloan Digital Sky Survey. Their high ultraviolet luminosities and lack of strong ultraviolet emission lines are suggestive of massive starbursts, although it is possible that they are more typical luminosity LBGs which have been highly magnified by strong gravitational lensing. Alternatively, they may be an unusual class of weak-lined quasars. If the ultraviolet and submillimetre properties of these objects mirror those of less luminous, starburst LBGs, then they should have detectable rest-frame far-infrared emission. However, our submm photometry fails to detect such emission, indicating that these systems are not merely scaled-up (either intrinsically or as a result of lensing) examples of typical LBGs. In addition we have searched for the morphological signatures of strong lensing, using high-resolution, near-infrared imaging, but we find none. Instead, near-infrared spectroscopy reveals that these systems are, in fact, a rare class of broad absorption-line (BAL) quasars. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, ?? -- ?? (2005)
Printed 23 January 2019
(MN LATEX style file v2.2)
Rest-frame optical and far-infrared observations of extremely bright
Lyman-break galaxy candidates at z ∼ 2.5
,
2 Ian Smail,3 Misty Bentz,4 J. A. Stevens,1
R. J. Ivison,1
S. C. Chapman6 and A. W. Blain6
1 UK Astronomy Technology Centre, Royal Observatory, Blackford Hill, Edinburgh EH9 3HJ
2 Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ
3 Institute for Computational Cosmology, University of Durham, South Road, Durham DH1 3LE
4 Department of Astronomy, The Ohio State University, 140 W. 18th Ave, Columbus, OH 43210-1173, USA
5 Centre for Astrophysics Research, Science and Technology Research Centre, University of Hertfordshire, College Lane, Hatfield AL10 9AB
6 Department of Astronomy, California Institute of Technology, MS 320-47, Pasadena, CA 91125, USA
,
5 K. Men´endez-Delmestre,6
23 January 2019
ABSTRACT
We have investigated the rest-frame optical and far-infrared properties of a sample of ex-
tremely bright candidate Lyman-break galaxies (LBG) identified in the Sloan Digital Sky
Survey. Their high ultraviolet luminosities and lack of strong ultraviolet emission lines are
suggestive of massive starbursts, although it is possible that they are more typical luminosity
LBGs which have been highly magnified by strong gravitational lensing. Alternatively, they
may be an unusual class of weak-lined quasars. If the ultraviolet and submillimetre proper-
ties of these objects mirror those of less luminous, starburst LBGs, then they should have
detectable rest-frame far-infrared emission. However, our submm photometry fails to detect
such emission, indicating that these systems are not merely scaled-up (either intrinsically or
as a result of lensing) examples of typical LBGs. In addition we have searched for the mor-
phological signatures of strong lensing, using high-resolution, near-infrared imaging, but we
find none. Instead, near-infrared spectroscopy reveals that these systems are, in fact, a rare
class of broad absorption-line (BAL) quasars.
Key words:
galaxies: evolution -- galaxies: formation -- galaxies: quasars -- galaxies:
active -- galaxies: individual: SDSS J024343.77−082109.9, SDSS J114756.00−025023.5,
SDSS J134026.44+634433.2, SDSS J143223.10−000116.4, SDSS J144424.55+013457.0,
SDSS J155359.96+005641.3.
1 INTRODUCTION
The very large areal coverage of the Sloan Digital Sky Survey
(SDSS, York et al. 2000) provides a unique opportunity to identify
rare, intrinsically luminous examples of high-redshift galaxy popu-
lations, as well as similarly rare, strongly-gravitationally magnified
examples of more typical high-redshift galaxies. To exploit this op-
portunity, Bentz & Osmer (2004) searched the SDSS Early Data
Release (Stoughton et al. 2002) for unusual quasars with anoma-
lously low C IV 154.9nm emission, and found a luminous z ∼ 2.5
starburst candidate that appeared to have been incorrectly classified
as a quasar by the SDSS pipeline.
Following on from this find, Bentz, Osmer & Weinberg
(2004) identified a further five sources from the SDSS First Data
Release (DR1) Quasar Catalog (Schneider et al. 2003) at z ∼ 2.5 --
2.8 with rest-frame ultraviolet (UV) colours similar to Lyman-
break galaxies (LBGs, Steidel et al. 2003) and exhibiting weak
or absent high-ionisation emission lines in their rest-frame UV
spectra. All six objects have r-band magnitudes of 19.8 -- 20.5,
with a median of r ∼ 20.3, i.e. they are an order of magnitude
brighter than the most luminous objects in existing LBG surveys.
Bentz, Osmer & Weinberg (2004) showed that if their UV emission
arises solely from star formation then they are intensely active star-
forming galaxies with strong lower limits on their star-formation
rates ranging from 300 to 1100 M⊙ yr−1, assuming negligible ab-
sorption by dust and adopting a continuous star-formation rate over
108 yrs (Kennicutt 1998). Thus these candidate LBGs could be
rare, extreme starbursts seen at the epoch, z ∼ 2.5, when both
the accretion luminosity density and the star-formation rate density
in the Universe are believed to peak (Miyaji, Hasinger & Schmidt
2000; Chapman et al. 2005).
Alternatively, the brightness and apparent rarity of these sys-
tems could simply reflect the fact that they are rare, strongly-
magnified examples of the normal-luminosity LBG population.
The LBG candidates show no evidence of multiple components at
the resolution of SDSS imaging (>
∼1 arcsec) and there are no obvi-
ous foreground lensing structures. Nevertheless, gravitational lens-
ing cannot yet be ruled out based on existing imaging.
2
Ivison et al.
The best argument against
these galaxies being highly-
magnified examples of normal LBGs results from their spectral
properties: the underlying continuua are much redder than typ-
ical LBGs, with observed 200 -- 600-nm spectral indices ranging
from α = −1.9 to −2.5 (where Fν ∝ ν +α), cf. α > −1.7
for most LBGs (Shapley et al. 2003), and the low-resolution SDSS
spectra show strong (sometimes broad) interstellar absorption lines
and Lyα emission in several cases, as well as hints of broad
C III] 190.9nm emission in two cases (Bentz, Osmer & Weinberg
2004). These features contrast with the narrow emission and in-
terstellar absorption lines seen in the composite LBG spectrum of
Shapley et al. (2003), but may be explained due to star-formation
activity (and perhaps associated winds) an order of magnitude
more vigorous than the sample considered by Shapley et al. (2003).
The gross spectral properties of a source should not be affected
markedly by lensing, so these differences argue that these galax-
ies are unlikely to be highly-magnified examples of the gen-
eral LBG population. Rather,
the spectral properties of these
sources share key characteristics with the submm-selected galax-
ies identified in the recent spectroscopic survey of Chapman et al.
(2003, 2005). In particular, they resemble N2 850.4, a compos-
ite starburst/AGN at z = 2.38 (Smail et al. 2003) and the BAL --
Sy2/QSO SMM J02399−0136 at z = 2.80 (Ivison et al. 1998;
Vernet & Cimatti 2001) -- the two-component Lyα emission,
broad C III] and the absorption seen in C II, Si IV, Al III and C IV
-- although the candidate LBGs lack such prominent P-Cygni pro-
files. The LBG candidates also share some common characteristics
with broad absorption-line (BAL) quasars -- the presence of broad
C III]190.9nm in two examples and some very broad absorption
lines -- although the line profiles are not typical of BALs. There
is thus a possibility that the UV emission from these galaxies is
powered by accretion rather than star formation, or that the sample
is a heterogeneous mix of star-forming galaxies and active galac-
tic nuclei (AGN). A recent paper by Hall et al. (2004) discusses
an unusual object, SDSS J113658.36+024220.1, whose rest-frame
UV spectrum shows a single emission line corresponding to Lyα
but no obvious metal-line emission, which bears some similari-
ties to the candidate LBGs studied here. Hall et al. (2004) inter-
pret SDSS J1136 as an AGN based in large part on tentative optical
variability and its strong radio emission, ∼1.4 mJy at 1.4 GHz.
If these galaxies are truly related to starburst LBGs then their
prodigious star formation should be betrayed in the rest-frame far-
infrared. Here, we exploit Submm Common-User Bolometer Ar-
ray -- SCUBA, Holland et al. (1999) -- submillimetre (submm)
photometry to search for such emission. We then use new, high-
resolution, near-infrared imaging to identify the morphological sig-
natures of strong lensing. Finally, we present near-infrared spectra
of these galaxies, covering a number of key rest-frame optical emis-
sion lines falling in the H and K atmospheric windows, to spectro-
scopically classify the galaxies and to search for quasar signatures
such as a tell-tale broad component to the Hα line. We describe our
observations in §2, present our results and discussion in §3 and give
our conclusions in §4.
with SCUBA on the James Clerk Maxwell Telescope (JCMT1).
On the first night, observations of SDSS J1147, SDSS J1340
and SDSS J1444 were made in average opacity conditions
(τ850µm ∼ 0.3 -- 0.4), and on the second night observations of
SDSS J0243 were made in better conditions (τ850µm ∼ 0.1 -- 0.2).
We used SCUBA in two-bolometer mode, giving a ∼15 per
cent improvement in signal-to-noise compared with one-bolometer
mode. Each source was observed for 1.8 ks. The STARLINK pack-
age SURF was used to reduce the data for each bolometer sepa-
rately. The resulting signals were then calibrated against the JCMT
secondary calibrators CRL 618 and 16293−2422, and co-added to
give weighted 850-µm flux densities and errors (see Table 1). Cal-
ibration uncertainties are estimated to be ∼10 per cent.
Near-infrared imaging data in the J- and K-bands were ob-
tained for the five bright LBG candidates accessible to the 3.8-m
UK Infrared Telescope (UKIRT2) during 2004 January -- April and
2004 July -- August. Flexible scheduling enabled us to utilise better-
than-average seeing on Mauna Kea, 0.4 -- 0.6 arcsec, and we em-
ployed the UKIRT Fast Track Imager, UFTI (Roche et al. 2003),
a 10242 HgCdTe array with 0.091-arcsec pixels, to exploit those
conditions. The total integration time in each filter, built up whilst
dithering every 60 s, was 3.8 ks. Contiguous observations of nearby
faint standards were used to determine zero points. The 3-σ de-
tection threshold is K ∼ 21.5 in a 4-arcsec-diameter aperture.
Data were reduced using ORAC-DR and we report effective total
magnitudes/colours (measured from 6-arcsec-diameter photome-
try) for the LBG candidates in Table 1. Objects in the frames were
then identified and catalogued using SExtractor (Bertin & Arnouts
1996) and colours measured in 2-arcsec-diameter apertures from
the aligned J- and K-band frames. K-band images of regions
around each target are shown in Fig. 1, with the (J −K) -- K colour-
magnitude distributions for each field displayed in Fig. 2.
Spectra were obtained during 2004 April and August with the
UKIRT 1 -- 5 µm Imager Spectrometer, UIST (Ramsay Howat et al.
1998), which utilises a 10242 InSb array with 0.12-arcsec pixels.
UIST's HK grism was used to cover the 1.4 -- 2.5 µm region, with
measured resolutions of λ/∆λ = 390 and 680 (for arc lines at 1.5
and 2.3 µm) for our 4-pixel-wide slit (0.48 × 120 arcsec). Acqui-
sition was accomplished using 20 -- 60-s sky-subtracted images of
each field. We are confident that all targets were placed within
a pixel of the optimal position on the slit. Each target was ob-
served for 6.7 ks, nodding along the slit in an A-B-B-A sequence
every 240 s. An Argon arc spectrum and a flatfield frame were
obtained prior to observations of each target. Nearby F5V stan-
dard stars were observed contiguously to set the flux scale, having
interpolated across their Hydrogen absorption lines before ratio-
ing. The frames were reduced using ORAC-DR, with optimal ex-
traction of spectra accomplished using FIGARO. For comparison
purposes we also obtained UIST HK-spectra of the weak-lined
quasar SDSS J113658.36+024220.1 (z = 2.4917, Hall et al. 2004)
and the z = 2.320 LoBAL quasar, SDSS J135317.80−000501.3
(Reichard et al. 2003; Willott et al. 2003). All the spectra are shown
in Fig. 3.
A further spectrum was obtained for the LBG candidate inac-
cessible to UKIRT (SDSS J134026.44+634433.2) using NIRSPEC
2 SUBMILLIMETRE AND NEAR-INFRARED
OBSERVATIONS
Submm photometry observations were obtained for four of our
LBG candidates in service time on 2004 January 15 and 28
1 The JCMT is operated by the Joint Astronomy Centre in Hilo, Hawaii, on
behalf of the Particle Physics and Astronomy Research Council (PPARC)
in the UK, the National Research Council of Canada, and The Netherlands
Organisation for Scientific Research.
2 UKIRT is operated by the Joint Astronomy Centre on behalf of PPARC.
Nature of extremely bright SDSS LBGs
3
Figure 1. Grey-scale representations of our deep K-band imaging of the five luminous SDSS LBG candidates. Note the cluster of faint K-band sources around
SDSS J1147 and the foreground group of bright galaxies near SDSS J1432. Each image is 60 × 60 arcsec with North to the top and East to the left and has
been smoothed by a Gaussian with a FWHM of 0.2 arcsec.
Figure 2. Distribution of sources on the (J − K) -- K colour-magnitude plane for each of our five target fields. We identify the candidate LBG in each field
by an ∗. We see a strong sequence in the distribution of galaxy colours in the field of SDSS J1432 with a typical colour of (J − K) ∼ 1.8, suggesting that
the over-density of galaxies in this field is at z ∼ 0.8. We plot only those galaxies with total K-band magnitudes brighter than the 5-σ limit of K = 21.0 and
show lower limits for those sources with <3-σ detections in our 2 arcsec-diameter photometric aperture in the J-band. The total area surveyed across the five
fields is 13.3 arcmin2.
on Keck-II3 during relatively poor conditions (0.9-arcsec seeing)
on 2005 March 18. The full spectral range was 2.27 -- 2.69 µm with
a resolution of ∼1,500, utilising a 42 × 0.76-arcsec (4-pixel) slit.
The total integration time was 1.6 ks, split into four A-B-B-A se-
quences with 100 s per exposure. A flux standard was not observed;
otherwise, the data reduction followed that employed for UIST.
3 RESULTS AND DISCUSSION
We now discuss the insights provided into the nature of the candi-
date LBGs by our various multi-wavelength observations. The ba-
sic properties of our sample -- redshifts, near-infrared photometry
and 850-µm flux densities -- are listed in Table 1. K-band imaging
of the target fields are shown in Fig. 1, (J − K) -- K colour magni-
tude diagrams of these fields in Fig. 2 and the HK spectra of the
candidate LBGs in Fig. 3. Line widths and fluxes for the stronger
features in these spectra are listed in Table 2.
3.1 Submillimetre properties
Studies of submm-selected galaxies (SMGs) have concluded that
most are too faint in the UV to be identified by the photometric
selection used for z ∼ 3 LBG surveys (Webb et al. 2003). This
suggests little overlap between these two well-studied classes of
high-redshift star-forming galaxies. However, statistical measure-
ments of the submm emission from samples of z ∼ 3 LBGs,
reaching below SCUBA's confusion limit, suggest that they may
contribute substantially to the SMG population at sub-mJy levels
3 The W. M. Keck Observatory is operated as a scientific partnership
among the California Institute of Technology, the University of California
and the National Aeronautics and Space Administration. The Observatory
was made possible by the generous financial support of the W. M. Keck
Foundation.
(Peacock et al. 2000; Webb et al. 2003; Kneib et al. 2004). Only a
handful of brighter examples are known (Chapman et al. 2002).
There is evidence of a more significant overlap between SMGs
and the UV-selected population identified at somewhat lower red-
shifts, z ∼ 1.5 -- 2.5 (Steidel et al. 2004), although again many of
the SMGs are too faint to be included in the photometric samples
(Chapman et al. 2005).
The raw star-formation rates estimated from the observed UV
luminosities of the SDSS sources, uncorrected for dust extinc-
∼ 1013 L⊙ for our sample, where SFR
tion, would imply LFIR >
= ǫ 10−10LFIR M⊙ yr−1 and ǫ = 0.8 -- 2.1 (Scoville & Young 1983;
Thronson & Telesco 1986), and thus 850-µm flux densities of 3 --
10 mJy (Blain & Longair 1996). Assuming a correction factor for
dust extinction typical of LBGs, 6× (Pettini et al. 2002; Erb et al.
2003), the predicted submm fluxes would increase by a similar fac-
tor. This suggests that the candidate LBGs should be detectable in
the submm waveband if they have submm/UV flux ratios similar to
the more typical luminosity members of this population. This con-
clusion holds whether these galaxies are either intrinsically bright
in the UV or are lensed (assuming that the lensing does not prefer-
entially boost the UV-bright regions).
The submm data presented in Table 1 demonstrate that the
four SDSS sources we have observed are all undetected individu-
ally at flux limits of 6 -- 8 mJy. This implies that it is unlikely that
these galaxies are simply scaled-up or strongly-lensed examples of
typical-luminosity z ∼ 3 LBGs. Indeed, the weighted mean for the
four sources (−0.36 ± 1.31 mJy) suggests that they would not have
been detected in even the deepest submm survey, although we can-
not rule out the possibility that the sample is heterogeneous, with a
handful of faint submm emitters.
For completeness we note that a search of the FIRST radio
survey (Becker, White & Helfand 1995) yielded only upper limits
at 1.4 GHz (Table 1), and that none of the galaxies were detected
by ROSAT to limits appropriate for the X-Ray All-Sky Survey.
4
Ivison et al.
Table 1. Near-infrared, submm and radio properties of the SDSS LBG candidates
Source name
z(UV)a
SDSS J024343.77−082109.9
SDSS J114756.00−025023.5
SDSS J134026.44+634433.2c
SDSS J143223.10−000116.4
SDSS J144424.55+013457.0
SDSS J155359.96+005641.3
2.590
2.556
2.786
2.472
2.670
2.635
K b
(mag)
17.33 ± 0.02
15.58 ± 0.02
16.90 ± 0.10
16.65 ± 0.03
17.59 ± 0.04
16.25 ± 0.03
(J − K)b
(mag)
1.62 ± 0.02
1.71 ± 0.03
1.00 ± 0.15
1.90 ± 0.04
1.07 ± 0.05
1.86 ± 0.04
(i − K)
(mag)
2.75
4.25
2.06
3.86
2.92
3.95
(r − K)
(mag)
3.07
3.71
2.76
3.47
2.49
3.43
S850µm
(mJy)
−2.13 ± 2.02
+2.61 ± 2.90
+1.93 ± 3.33
...
−1.33 ± 2.78
...
S1.4GHz
(mJy)
5σ < 0.97
5σ < 1.03
5σ < 0.98
5σ < 0.97
5σ < 1.01
5σ < 0.94
a UV redshifts were measured using cross-correlation with a quasar template and searches for emission lines (Schneider et al. 2002).
b Aperture magnitudes, measured in 6-arcsec-diameter apertures. Correction for line contamination is discussed in §3.2.
c Near-infrared photometry from Teplitz et al. (2004).
Table 2. Spectral properties of the SDSS LBG candidates and comparison quasars
Source
Name
SDSS J024343.77−082109.9
SDSS J114756.00−025023.5
SDSS J134026.44+634433.2
SDSS J143223.10−000116.4
SDSS J144424.55+013457.0
SDSS J155359.96+005641.3
SDSS J113658.36+024220.1
SDSS J135317.80−000501.3
FWHM(Hα)
(km s−1)
Flux(Hα)
(10−16 W m−2)
Flux([OIII]5007)
(10−16 W m−2)
5,360
13,830
12,200b
6,150
5,730c
7,190
4,100
5,180
1.4 ± 0.1
5.3 ± 0.2
...
8.6 ± 0.2
2.1 ± 0.3c
11.6 ± 0.3
3.5 ± 0.1
6.1 ± 0.2
3σ < 0.3
0.5 ± 0.1
...
1.3 ± 0.1
0.6 ± 0.1
0.5 ± 0.1
0.3 ± 0.1
3σ < 6.0
Flux(Hβ)
EWa (Hα)
z([OIII])
z(Hα)
(10−16 W m−2)
3σ < 0.3
0.4 ± 0.1
...
2.1 ± 0.1
3σ < 0.4
2.0 ± 0.2
0.4 ± 0.1
3σ < 0.3
(nm)
−96
−152
−164b
−166
−30c
−157
−106
−139
2.5940
2.5701
...
2.4728
2.6767
2.6350
2.4928
--
2.5995
2.5668
...
2.4772
2.6715
2.6404
2.4946
2.3182
a Not corrected to the rest frame.
b Fits assume z(Hα) = z(UV) = 2.786.
c Lower limits, given the lack of continuum redward of the line.
3.2 Photometric and morphological properties
Our near-infrared observations indicate that the median observed
colours for the LBG candidates are J − K = 1.71 ± 0.20,
i − K = 3.86 ± 0.43 and r − K = 3.43 ± 0.37. Comparing
the J − K colours with the spectra of the sources, from §3.3, it
is clear that the LBG candidates with the strongest line emission
in the K-band also have the reddest continuum colours, suggest-
ing that the line emission is biasing the colours we measure. With
typical observed-frame equivalent widths of −100 to −150 nm,
and a K-band filter width of ∼350 nm, the fluxes in K should be
corrected by approximately 0.6 -- 0.7× (or +0.5 in magnitudes). No
sources are classed as extremely red objects on the basis of either
their r − K or i − K colours, although the reddest source in the
optical/near-infrared, SDSS J1147, is also the reddest in the rest-
frame UV and is the only candidate LBG in our sample which does
not show Lyα emission. The median J − K colour of our sam-
ple is comparable to that seen for z ∼ 3 LBGs from Shapley et al.
(2001), J − K ∼ 1.63, although our candidate LBGs are redder
on average in r − K than the standard UV-selected populations at
z ∼ 2 or z ∼ 3, r − K ∼ 3.25 and r − K ∼ 2.85, respec-
tively (Shapley et al. 2001; Steidel et al. 2004). This suggests that
the rest-frame UV continua may be significantly redder than nor-
mal LBGs, although their rest-frame optical continua are compara-
ble (before correcting for emission-line contributions).
Turning to the high-resolution near-infrared imaging (Fig. 1),
we find that all five candidate LBGs are unresolved at the 0.45 --
0.55-arcsec seeing of our K-band images (as measured from mul-
tiple stars in each frame). Given the signal to noise of our detec-
tions of the LBGs and their measured FWHM relative to stars in
the fields, we can place firm limits of < 0.1 arcsec FWHM on the
sizes of these sources or on the separation of multiple components
if they are strongly lensed. In the absence of lensing, this angu-
lar limit corresponds to an intrinsic size of <1 kpc for the physical
scale of these sources in their rest-frame V -band light.
We identify none of the morphological signatures expected
from strong galaxy -- galaxy lensing in our deep K-band images:
multiple lensed components or an identifiable foreground lens. We
note that the angular size limit estimated above, if taken as the Ein-
stein diameter, would correspond to a velocity dispersion of only
50 km s−1 (for a spherical isothermal lens at z ∼ 0.5). This ve-
locity dispersion would correspond to a ∼0.1 L∗ early-type galaxy
with an expected magnitude of K ∼ 19 at z ∼ 0.5, detectable in
our imaging out to z ∼ 1 (Rusin et al. 2003). No such nearby lenses
are visible in our imaging on the relevant scales (0.2 -- 2 arcsec).
Looking at a wider region around the candidate LBGs we see
a compact K = 17.9 galaxy (with J − K ∼ 0.8) lying only 4 arcsec
away from SDSS J1553, but this would not provide a strong am-
plification of the source. We also find that SDSS J1432 sits in the
outskirts of a dense, compact foreground group, the brightest mem-
bers of which have K ∼ 16. It is possible that the LBG candidate
is thus magnified by weak lensing from the foreground structure,
although the total magnification is likely to be modest. The fields
surrounding SDSS J0243 and SDSS J1444 are unremarkable, but
we do identify a group of 5 -- 6 faint, resolved galaxies, K >
∼ 19.4,
within 15 arcsec of SDSS J1147. These are possibly members of a
cluster, either in the foreground or (given their faintness) associated
with the candidate LBG. These faint galaxies exhibit a wide range
in J −K colours, 0.8 -- 3.0, including some extremely red objects. In
addition, several brighter galaxies, K >
∼ 17.3, lie within 20 arcsec.
Nature of extremely bright SDSS LBGs
5
Figure 3. The six spectra at the bottom of this panel illustrate the rest-frame optical spectra of the six luminous SDSS LBG candidates (arbitrarily offset in
flux). SDSS J1340, observed with NIRSPEC, has only K-band coverage. Above these, for comparison, are similar spectra of the LoBAL QSO, SDSS J1353,
and the Lyα-only AGN, SDSS J1136. All the spectra are smoothed with a Gaussian of σ = 0.75 pixels. We mark the wavelengths of possible absorption or
emission lines which may be visible in these galaxies.
We cannot demonstrate a significant over-density, but they could be
more luminous members of the same structure.
Looking at the (J − K) -- K colour-magnitude plots (Fig. 2)
for the five fields we see that the LBG candidates are rarely the red-
dest galaxy in the field, although both SDSS J1147 and SDSS J1432
have very red, close neighbours (J − K = 2.2 -- 3.0). The most
striking feature is the colour-magnitude sequence seen in the distri-
bution of galaxy colours in the SDSS J1432 field. This has a char-
acteristic colour of J − K ∼ 1.8 at K ∼ 17.5, consistent with that
expected for evolved galaxies in a group or poor cluster at z ∼ 0.7 --
0.8 (Feulner et al. 2003). The close similarity of the colours of the
candidate LBG to these galaxies may either be a coincidence or
could point to contamination of our photometric measurement by a
superimposed member of this group (which could thus also gravita-
tionally magnify the background source). We believe that the sim-
ilarities in the colours are merely a coincidence as it is clear from
the spectrum of SDSS J1432 (Fig. 3) that there is a significant con-
tribution to the K-band light from the background source.
We conclude that the morphological information for all five
of the candidate LBGs we have imaged at <∼0.55-arcsec resolution
provides no support for them being strongly lensed. There is pos-
sible evidence for weak lensing for one or more sources, but this
would not significantly affect the apparent magnitudes of these sys-
tems.
3.3 Spectral properties
It is immediately apparent from the rest-frame optical spectra pre-
sented in Fig. 3 that all six of the spectroscopically-observed LBG
candidates contain broad-line AGN. All have broad Hα emission,
6
Ivison et al.
with FWHM line widths ranging from 5,000 to ∼14,000 km s−1
(Table 2) and broad components visible in Hβ for several sys-
tems (despite the typically poorer signal-to-noise in the H-band).
SDSS J1444 has the weakest emission and there is no spectral cov-
erage beyond the red extreme of the line due to its high redshift, but
a broad Hα line is still apparent. Comparing them to the two known
AGN we observed, the weak-lined AGN SDSS J1136 and the low-
ionisation BALQSO SDSS J1353, we see that the candidate LBGs
exhibit broader Hα emission than either of these two AGN.
So, clear AGN features are visible in the rest-frame optical,
whereas the UV spectra of these galaxies are characterised by
a strong continuum but lack the strong emission lines typical of
AGN. Does the AGN contribute significantly to the UV fluxes of
these galaxies?
The redshift measurements available to us tend to follow
the same pattern for all the LBG candidates: the UV-determined
values, based predominantly on the prominent Lyα emission
line, are slightly blueward of the [O III] 500.7nm and Hα lines
(∆z = 0.0051 ± 0.0057 and 0.0065 ± 0.0037, respectively). The
only exception is SDSS J1444, where the Lyα and Hα redshifts are
identical, but the Hα redshift is poorly determined. The Hα emis-
sion we see originates close to the AGN, in the broad-line region
(BLR), so it is natural to assume that the UV absorption lines are
due to wind-driven material in our line of sight to the BLR, and
that the UV continuum also arises close to the AGN. Indeed, there
is a tight correlation between the FWHM of the Hα (Table 2) and
the absolute i-band magnitudes from Bentz, Osmer & Weinberg
(2004): the 0.07-dex scatter suggests a close relationship between
the UV continuum emission and the AGN. However, it is not clear
whether this is a direct relationship, or whether it arises merely be-
cause more massive AGN reside in more luminous galaxies. Nev-
ertheless, assuming that the candidate LBGs have intrinsic power-
law continua characteristic of normal quasars, with α = −0.44
(Vanden Berk et al. 2001), then their observed rest-frame 200 -- 600-
nm spectral slopes (α = −1.89 to −2.54) indicate substantial
dust extinction, AV ∼ 1.35 -- 1.95, for a Calzetti extinction law
(Calzetti et al. 2000).
But where are the UV emission lines usually associated with
quasar activity? If they have been quenched by dust surrounding
the active nuclei, why can we still see intense UV continuum emis-
sion? The spectral characteristics of these galaxies are unusual,
but the presence of strong and broad absorption lines in the UV
are similar to those seen in less-reddened examples of the most
extreme BAL quasars found by the SDSS (Hall et al. 2002) and
in the Digitized Palomar Observatory Sky Survey (Brunner et al.
2003). Indeed, very recently Appenzeller et al. (2005) have pub-
lished high-resolution ´echelle spectroscopy of SDSS J1553 which
provides much higher-quality information about the UV spectral
properties of this galaxy. Based on their analysis of the detailed
properties of the absorption lines, they conclude that SDSS J1553
is a low-ionisation BAL quasar (LoBALQSO), or perhaps an even
rarer FeLoBALQSO. The very strong low-ionisation absorption
features found in LoBALQSOs across a wide velocity range can
strongly suppress the emission-line components in these systems,
leading to the absorption-dominanted UV spectra we see. The UV
absorption features of SDSS J1553 are typical of those seen in the
other five galaxies and so we expect that deeper and higher res-
olution spectroscopy of the complete sample would likely lead to
the same conclusion for the other sources. Indeed, based on the ex-
isting low-resolution SDSS spectra, Appenzeller et al. (2005) sug-
gest at least two-thirds of the sample may be LoBALQSOs. The
relative weakness and narrowness of the BAL features, combined
with the absence of strong UV emission features, suggests that
the outflows in these galaxies may differ in terms of their veloc-
ity and spatial coverage compared to those seen in typical LoB-
ALQSOs. Alternatively, these AGN may be similar to SDSS J1136
(Fig. 3), which Hall et al. (2004) suggest for some unknown reason
has weak, broad and highly blue-shifted emission lines.
Finally, we want to highlight the properties of SDSS J1340.
This candidate LBG was detected at 16 µm using Spitzer by
Teplitz et al. (2004). They interpret this detection in terms of a mas-
sive starburst, even though the optical/mid-infrared spectral energy
distribution (SED) of SDSS J1340 in Teplitz et al. (2004) is best fit
by the SED for the Seyfert-1, NGC 5548. Our non-detection of the
source in the submm suggests that the 16 µm detection most likely
arises from high-temperature AGN-heated dust, rather than a bolo-
metrically luminous starburst. Further support for the presence of
a bolometrically luminous AGN in this system comes from the de-
tection of a strong and broad Hα line in our near-infrared spectrum
(Fig. 3).
4 CONCLUSIONS
We present multi-wavelength observations of a sample of six candi-
date LBGs at z = 2.5 -- 2.8 identified from the SDSS DR1 QSO Cat-
alog by Bentz, Osmer & Weinberg (2004). We suggest that these
sources could be either: 1) intrinsically luminous, UV-bright star-
bursts; 2) strongly-lensed examples of typical-luminosity LBGs; or
3) a class of quasars with extremely weak UV emission lines.
We do not detect any of the four candidate LBGs observed
in the submm, placing a strong constraint on the submm emission
from the ensemble. This suggests that the sources are unlikely to be
strongly-lensed examples of more typical LBGs, or intrinsically-
luminous LBGs, unless the far-infrared emission from such UV
starbursts declines precipitously at high luminosities. Two further
pieces of evidence weigh against the lensing hypothesis: first, the
UV spectral properties of the candidate LBGs do not match those
of typical luminosity LBGs; second, using high-resolution near-
infrared imaging of five of the candidates we find no morpholog-
ical evidence of strong lensing. Taking these results together, we
conclude that the sources in our sample are unlikely to be either
intrinsically-luminous LBGs or rare, strongly-lensed examples of
more normal LBGs. This suggests that they are most likely to be
unusual AGN.
Our near-infrared spectroscopy confirms this suggestion,
identifying very broad lines in the rest-frame optical spec-
tra of all six galaxies in the sample. We therefore conclude
that the six apparently extremely luminous LBGs identified by
Bentz, Osmer & Weinberg (2004) are likely to be LoBALQSOs
whose unusually weak UV emission lines may either be an intrinsic
property of these AGN (Hall et al. 2004) or result from a complex
distribution of absorption in the outflow close to the AGN.
ACKNOWLEDGMENTS
We thank Pat Osmer and David Weinberg for their work on
the SDSS LBG survey. We also thank Alastair Edge for useful
conversations, and the referee for suggestions that improved the
paper markedly. We acknowledge service observations from the
JCMT. IRS acknowledges support from the Royal Society. MB
is supported by a Graduate Fellowship from the National Science
Foundation. AWB acknowledges support from NSF grant AST-
0205937, the Research Corporation and the Alfred P. Sloan Foun-
dation.
Willott C.J., Rawlings S., Grimes J.A., 2003, ApJ, 598, 909
York D.G. et al. 2000, AJ, 120, 1579
Nature of extremely bright SDSS LBGs
7
REFERENCES
Appenzeller I., Stahl O., Tapken C., Mehlert D., Noll S., 2005,
A&A, 435, 465
Becker R.H., White R.L., Helfand D.J., 1995, ApJ, 450, 559
Bertin E., Arnouts S., 1996, A&AS, 117, 393
Bentz M.C., Osmer P.S., 2004, AJ, 127, 576
Bentz M.C., Osmer P.S., Weinberg D.H., 2004, ApJ, 600, L19
Blain A.W., Longair M.S., 1996, MNRAS, 279, 847
Brunner R.J. et al. 2003, AJ, 126, 53
Calzetti D., Armus L., Bohlin R.C., Kinney A.L., Koornneef J.,
Storchi-Bergmann T., 2000, ApJ, 533, 682
Chapman S.C., Shapley A., Steidel C., Windhorst R., 2002, ApJ,
572, L1
Chapman S.C., Blain A.W., Ivison R.J., Smail I. 2003, Nature,
422, 695
Chapman S.C., Blain A.W., Smail I., Ivison R.J. 2005, ApJ, 622,
772
Erb D. et al. 2003, ApJ, 591, 101
Feulner G., Bender R., Drory N., Hopp U., Snigula J., Hill G.J.,
2003, MNRAS, 342, 605
Hall P.B. et al. 2002, ApJS, 141, 267
Hall P.B. et al. 2004, AJ, 127, 3146
Holland W.S. et al. 1999, MNRAS, 303, 659
Ivison R.J., Smail I., Le Borgne J.-F., Blain A.W., Kneib J.-P.,
B´ezecourt J., Kerr T.H., Davies J.K. 1998, MNRAS, 298, 583
Kennicutt R.C., 1998, ARA&A, 36, 189
Kneib J.-P., van der Werf P.P., Kraiberg Knudsen K., Smail I.,
Blain A., Frayer D., Barnard V., Ivison R., 2004, MNRAS, 349,
1211
Miyaji T., Hasinger G., Schmidt M., 2000, A&A, 353, 25
Peacock J.A. et al., 2000, MNRAS, 318, 535
Pettini M., Rix S.A., Steidel C.C., Adelberger K.L., Hunt M.P.,
Shapley A.E., ApJ, 2002, 569, 742
Ramsay Howat S.K. et al., 1998, SPIE, 3354, 456
Reichard T.A. et al., 2003, AJ, 125, 1711
Roche P.F. et al., 2003, SPIE, 4841, 901
Rusin D. et al., 2003, ApJ, 587, 143
Schneider D.P. et al., 2002, AJ, 123, 567
Schneider D.P. et al., 2003, AJ, 126, 2579
Scoville N.Z., Young J.S., 1983, ApJ, 265, 148
Shapley A.E., Steidel C.C., Adelberger K.L., Dickinson M., Gi-
avalisco M., Pettini M., 2001, ApJ, 562, 95
Shapley A.E., Steidel C.C., Pettini M., Adelberger K.L., 2003,
ApJ, 588, 65
Smail I., Chapman S.C., Ivison R.J., Blain A.W., Takata T., Heck-
man T.M., Dunlop J.S., Sekiguchi K., 2003, MNRAS, 342, 1185
Steidel C.C., Adelberger K.L., Shapley A.E., Pettini M., Dickin-
son M., Giavalisco M., 2003, ApJ, 592, 728
Steidel C.C., Shapley A.E., Pettini M., Adelberger K.L., Erb D.K.,
Reddy N.A., Hunt M.P., 2004, ApJ, 604, 534
Stoughton C. et al., 2002, AJ, 123, 485
Teplitz H.I. et al., 2004, ApJS, 154, 103
Thronson H., Telesco C., 1986, ApJ, 311, 98
Vanden Berk D.E. et al., 2001, AJ, 122, 549
Vernet J., Cimatti A., 2001, A&A, 380, 409
Webb T.M.A. et al. 2003, ApJ, 582, 6
|
astro-ph/9906384 | 1 | 9906 | 1999-06-24T14:51:07 | Milagrito Detection of TeV Emission from Mrk 501 | [
"astro-ph"
] | The Milagro water Cherenkov detector near Los Alamos, New Mexico, has been operated as a sky monitor at energies of a few TeV between February 1997 and April 1998. Serving as a test run for the full Milagro detector, Milagrito has taken data during the strong and long-lasting 1997 flare of Mrk 501. We present results from the analysis of Mrk 501 and compare the excess and background rates with expectations from the detector simulations. | astro-ph | astro-ph | OG2.1.11
1
Milagrito Detection of TeV Emission from Mrk 501
Stefan Westerhoff for the Milagro Collaboration
University of California, Santa Cruz, CA 95064, USA
Abstract
The Milagro water Cherenkov detector near Los Alamos, New Mexico, has been operated as a sky monitor
at energies of a few TeV between February 1997 and April 1998. Serving as a test run for the full Milagro
detector, Milagrito has taken data during the strong and long-lasting 1997 flare of Mrk 501. We present results
from the analysis of Mrk 501 and compare excess and background rate with expectations from the detector
simulation.
1 Introduction:
With the detection of 4 Galactic and 3 extragalatic sources, Very High Energy (VHE) (cid:13) -ray astronomy,
studying the sky at energies above 100 GeV, has become one of the most interesting frontiers in astronomy.
Source detections and analyses in this field are still dominated by the highly successful atmospheric Cherenkov
technique. Cherenkov telescopes and telescope arrays are optimal tools for the detailed study of established
sources and their energy spectra and the theory-guided search for yet unknown sources. There is, however,
also a strong case for instruments able to perform an unbiased, systematic and continuous search for TeV
sources, thus overcoming the limitations imposed by the low duty cycle and small field of view of Cherenkov
telecopes. Consequently, the observation technique must exploit the particle content of air showers rather than
the Cherenkov light.
A first-generation all-sky monitor operating at energies below 1 TeV, the Milagro detector (McCullough
et al., 1999) located 2650 m above sea level near Los Alamos, New Mexico, at latitude (cid:21) = :
o N, started
data taking in early 1999. Milagro is a water Cherenkov detector of size (cid:2) (cid:2) m
. Two layers of
photomultiplier tubes detect the Cherenkov light produced by secondary particles entering the water. The
first layer, with 450 tubes on a (cid:2) m
grid at a depth of 1.4 m, allows the shower direction and thus the
direction of the primary particle to be reconstructed, while the second layer with 273 tubes at a depth of
(cid:24)7.0 m primarily detects the penetrating component of air showers, i.e. muons, hadrons, and highly energetic
electromagnetic particles.
A smaller, less sensitive prototype, Milagrito (Atkins et al., 1999), has taken data between February 1997
and April 1998. Milagrito, a one-layer detector of size (cid:2) (cid:2) m
with 228 photomultiplier tubes on a
grid at a rather shallow depth of 0.9 m, served mainly as a test run for this relatively new detection
(cid:2) m
technique. This prototype has, however, taken data during a very intense and long-lasting flare of Mrk 501 in
1997 (Samuelson et al., 1998).
For the evaluation of the performance of VHE instruments, the Crab nebula is usually used as a “standard
candle”. It is a well-studied steady source with a flux of
J(cid:13) (E ) = (: (cid:6) : (cid:6) :) (cid:2)
(cid:0)
(cid:0): (cid:6) : (cid:6) :
E
TeV
(cid:0)
(cid:0)
s
m
TeV
(cid:0)
;
(1)
(Hillas et al., 1998). Simulations indicate that the expected significance from Milagrito for the Crab nebula is
less than (cid:27) , ruling out the possibility of using a Crab signal to test Milagrito’s performance. A detection of
Mrk 501 with a sufficiently high significance can be expected had the average flux been in excess of the Crab
flux. During its flare in 1997, Mrk 501 has been intensively studied with several air Cherenkov telescopes.
Although not covering the same observation times, the average fluxes measured by Whipple (Samuelson et
al., 1998) and the HEGRA stereo system of air Cherenkov telescopes (Aharonian et al., 1999) agree extremely
well both in shape and magnitude, and they both indicate a significant deviation of the energy spectrum from
Figure 1: (a) Effective area of Milagrito for reconstructed (cid:13) - and cosmic ray showers, averaged over a zenith
angle range from
o, as a function of the primary energy. (b) Effective area for (cid:13) -showers for
o
(cid:20) (cid:18) (cid:20)
various zenith angle ranges.
a simple power law. Using an average fl ux as measured by Whipple,
J(cid:13) (E ) = (: (cid:6) : (cid:6) :) (cid:2)
(cid:0)
(cid:0): (cid:6) : (cid:6) : (cid:0)( : (cid:6) : ) log E
E
TeV
(cid:0)
(cid:0)
s
m
TeV
(cid:0)
;
(2)
simulations of the Milagrito detector response predict the expected integral (cid:13) -rate from Mrk 501 to be 3.6 times
the Crab rate. Although highly variable sources like Mrk 501 are not well-suited for checking the sensitivity
of detectors integrating over long time periods, the observation of an excess from Mrk 501 still provides a test
for the sensitivity of Milagrito and reliability of the detector simulation.
In addition, observations with Cherenkov telescopes cover only the time from February to October, while
Milagrito continued to monitor Mrk 501 in late 1997 and early 1998.
2 Milagrito Performance:
Sensitivity predictions for Milagrito are based on a detector simulation using the CORSIKA 5.61 air shower
simulation code for the development of the shower in the Earth ’ s atmosphere, and the GEANT 3.21 package
for the simulation of the detector. The simulation is described in detail elsewhere (Atkins et al.,1999).
The Milagrito detector operated with a minimum requirement of 100 hit tubes per event. Figure 1 (a) shows
the effective area Ae(cid:11) of Milagrito for (cid:13) -showers and cosmic ray background showers induced by protons,
helium, and nitrogen, the latter used for representing the combined CNO fl ux, as a function of the energy of
the primary particle. Figure 1 (b) shows how the effi ciency depends on the zenith angle (cid:18).
At energies (cid:20) TeV, the effective area for proton-induced showers is larger than for (cid:13) -showers. This
is related to the fact that (cid:13) -induced (thus almost purely electromagnetic) showers are usually more laterally
con fi ned so the area covered by the particles reaching detector altitude is smaller than for hadron-induced
showers, which tend to have “ hot spots ” with high particle density at large distances from the shower core.
At energies above (cid:24) 5 TeV, the larger effective area for (cid:13) -induced showers provides an intrinsic cosmic ray
background rejection.
Figure 2: (a) Angular resolution and (b) energy distribution of (cid:13) -showers triggering Milagrito.
As a large fraction of showers fulfi lling the trigger condition have their core outside the sensitive detector
area, the effective area is larger than the geometrical area above (cid:24) 3 TeV. In fact, only % of the proton
showers and % of the (cid:13) -showers triggering the Milagrito detector have their core on the pond. This leads
to a rather broad energy distribution starting at energies as low as 100 GeV, with no well defi ned threshold
energy (Figure 2 (b)). The median energy varies slightly with the source declination (cid:14) , ranging from (cid:24) 3 TeV
o to 7 TeV for sources with j(cid:14) (cid:0) (cid:21)j ’
o.
for sources at (cid:14) =
The water Cherenkov technique uses water both as the converter and the detector medium. Consequently,
the effi ciency for detecting low energy air shower particles is very high, leading to a good sensitivity even
tting, however, has to deal with a considerable
for showers with primary energy below 1 TeV. The angle fi
amount of light late as compared to the shower front reaching the detector. The “ late light”
is partly produced
by low energy particles which tend to trail the shower front. More important, however, is the horizontal
light component resulting from the large Cherenkov angle in water (
o), multiple scattering, (cid:14) -rays, and
scattering and refl ection of Cherenkov light. The expected angular resolution for cosmic rays agrees with our
observations of the cosmic ray shadow of the moon (Wascko et al., 1999).
t to the arrival times
Milagrito’ s angular resolution is a strong function of the number of the tubes in the fi
of the tubes (Figure 2 (a)). For the initial source search, a minimum number of 40 tubes used in the shower
t is required. This leads to a measured rate of (cid:6) reconstructed events per day from cosmic ray
plane fi
showers in a typical source bin with :
o radius at the declination of Mrk 501. This is in good agreement with
(cid:0) events per day from protons, Helium, and CNO nuclei. In the simulation, the
the predicted rate of
+
contribution of He and CNO to the total trigger rate turns out to be % and %, respectively.
3 Results:
A straight-forward analysis with a source bin of radius :
o centered on Mrk 501 leads to an excess > (cid:27) .
According to simulations this bin size contains % of the source events and is optimal for an analysis treating
all events equally. The corresponding excess rate averaged over the lifetime of Milagrito (370 equivalent
source days for Mrk 501) is (: (cid:6) : ) day
(cid:0) .
Figure 3: Excess/background for Mrk 501 as a function of time. At the current sensitivity level the data is
consistent with a constant fl ux.
Figure 3 shows how the excess is accumulated over Milagrito’ s lifetime. At our present level of sensitivity,
the data is consistent with a fl ux constant in time.
Using the average fl ux as measured by air Cherenkov telescopes between February and October 1997,
simulations predict a (cid:13) -rate of (: (cid:6) : ) day
(cid:0) for Milagrito and are thus consistent with the measured
excess during this period, (: (cid:6) :) day
(cid:0) .
An analysis that takes account of the strong dependence of the resolution on the number of photomultipliers
in the fi
t should be more sensitive to emission from a point source. The results of such an analysis will be
presented at the conference.
The analysis was extended to 10 other nearby blazars (z (cid:20) : ) in Milagrito’ s fi eld of view, but Mrk 501
remains the only analyzed source with a signifi cance in excess of (cid:27) . Results from this blazar sample are
reported elsewhere (Westerhoff et al., 1999).
This research was supported in part by the National Science Foundation, the U. S. Department of Energy
Offi ce of High Energy Physics, the U. S. Department of Energy Offi ce of Nuclear Physics, Los Alamos
National Laboratory, the University of California, the Institute of Geophysics and Planetary Physics, The
Research Corporation, and the California Space Institute.
References
Aharonian, F. et al. 1999, Astron. & Astrophys., in press.
Atkins et al. 1999, Nucl. Instr. Meth. A, in preparation.
McCullough, J.F. et al., these ICRC proceedings (HE 6.1.02).
Hillas, A.M. et al. 1998, ApJ 503, 744.
Samuelson, F.W. et al., 1998, ApJ 501, L17.
Wascko, M.O. et al., these ICRC proceedings (SH 3.2.39).
Westerhoff, S. et al. 1999, Proc.
th Texas Symposium on Relativistic Astrophys., Paris (France), 1998.
|
0801.1372 | 1 | 0801 | 2008-01-09T08:49:34 | Extragalactic H3O+: Some Consequences | [
"astro-ph"
] | We discuss some implications of our recent detection of extragalactic H3O+: the location of the gas in M82, the origin of energetic radiation in M82, and the possible feedback effects of star formation on the cosmic ray flux in galaxies. | astro-ph | astro-ph | Far-infrared observations of the interstellar medium
Editors: C. Kramer, R. Simon et al.
EAS Publications Series, Vol. ?, 2008
8
0
0
2
n
a
J
9
]
h
p
-
o
r
t
s
a
[
1
v
2
7
3
1
.
1
0
8
0
:
v
i
X
r
a
EXTRAGALACTIC H3O+: SOME CONSEQUENCES
Floris van der Tak 1, Susanne Aalto 2 and Rowin Meijerink 3
Abstract. We discuss some implications of our recent detection of ex-
tragalactic H3O+: the location of the gas in M82, the origin of energetic
radiation in M82, and the possible feedback effects of star formation
on the cosmic ray flux in galaxies.
1
Introduction
Last year saw the first detection of the H3O+ molecule outside the Galaxy (Van der Tak et al.
2008). Using the new 16-pixel HARP imaging spectrometer on the James Clerk
Maxwell Telescope, line emission at a rest frequency of 364 GHz was detected to-
wards the prototypical starburst galaxy M82 and the prototypical ultraluminous
merger Arp 220. The derived H3O+ abundances and H3O+/H2O ratios imply
very high ionization rates for the dense molecular gas in the nuclei of these two
galaxies. Chemical models (Meijerink et al. 2007) indicate that the origin of this
high ionization is irradiation by X-rays in the case of Arp 220, whereas for M82,
a combination of ultraviolet light and cosmic rays is needed. The high ionization
rates of these galactic nuclei make magnetic fields more effective in retarding their
gravitational collapse and the formation of stars. The origin of the irradiation in
M82 is the evolved starburst (Forster Schreiber et al. 2003), whereas for Arp 220,
an AGN may be needed, as claimed before by Downes & Eckart (2007). Naturally,
some questions remain, of which this paper discusses a few.
2 Location of the H3O+ in the nucleus of M82
The velocities and widths of the two components of the H3O+ line profile observed
toward M82 may be used to constrain the location of the gas. Observations of CO
emission lines toward the M82 nucleus show a flattened structure of size ≈50×20′′
1 SRON, Landleven 12, 9747 AD Groningen, The Netherlands; [email protected]
2 Onsala Space Observatory, Sweden
3 University of California at Berkeley, USA
c(cid:13) EDP Sciences 2018
DOI: (will be inserted later)
2
Far-infrared observations of the interstellar medium
(1000×400 pc) with emission peaks on its north-eastern and south-western ends.
The velocity field is well described by a monotonic gradient along the major axis,
so that the structure is probably a rotating ring or torus. Observations of CO 6 -- 5
with the JCMT (Seaquist et al. 2006) show a third peak in the middle, which is
even better visible in interferometric images of CO 1 -- 0 (Walter et al. 2002) and
known as the 'central hotspot'.
The simple velocity structure of the rotating ring implies that the velocities of
the H3O+ components correspond to unique positions. In particular, the velocity
of the narrow component of Vhelio≈270 km s−1 corresponds to the inner edge of the
north-eastern peak, which is just covered by the beam of the H3O+ observations.
The width of the narrow component agrees with that of the CO emission from the
north-eastern peak.
The velocity of the broad H3O+ component of Vhelio≈220 km s−1 corresponds
to a position right between the two main CO peaks. Since the width of this
component is much larger than the systematic velocity gradient within one JCMT
beam, the most plausible origin of the broad component is the 'central hotspot' of
molecular emission. This hotspot is located close to the dynamical center of M82
and shows large line widths in several other molecular lines. The distribution of
H3O+ in M82 thus appears to mimic that of other dense gas tracers such as HCN
(Mauersberger & Henkel 1991).
3 Origins of energetic radiation in M82
The X-ray luminosity (LX ) of M82 on arcminute scales as measured by ROSAT
and ASCA is ∼2×1041 erg/s after correction for internal absorption (Moran & Lehnert
1997). In the central < sim100¡,pc of M82, where H3O+ has been detected, both
the nuclear non-thermal component and the central thermal component may con-
tribute to the ionization of the molecular gas; the superwind is too far away and
its luminosity too low for it to play a role. Observations with Chandra and XMM-
Newton (Zezas 2006) seem to indicate that the emission in the extended nuclear
region is dominated by high-mass X-ray binaries (HMXBs) and so-called ultra-
luminous X-ray sources (ULXs). The X-ray emission from a starburst is domi-
nated by HMXBs, such that LX may be used to estimate its star formation rate
(Grimm et al. 2003). Future spatially and spectrally resolved observations of H2O
line emission in M82 may be used to determine the H3O+/H2O ratio for both com-
ponents of H3O+ emission and to characterize the ionization state of the molecular
gas in M82 as a function of position in the nuclear disk.
4 Effects of star formation on the cosmic-ray flux
The main astrophysical interest in H3O+ is the use of the H3O+/H2O abundance
ratio as a probe of the ionization rate of dense molecular gas. This rate is an
important parameter for the ability to form stars, because it determines whether
magnetic fields may be effective in supporting the cloud against gravitational col-
lapse. The ionization rate of dense molecular clouds is dominated by ultraviolet
Floris van der Tak et al.: Extragalactic H3O+: Some Consequences
3
light and X-rays near their surfaces, and low-energy cosmic rays (E <
∼ 1 GeV) in
their interiors. Foreground absorption often hides these types of radiation from
direct observation, which is why H3O+ observations are so useful.
Recently, Pellegrini et al. (2007) proposed a model for the photodissociation
region M17 where the thickness of the atomic gas layer between the ionized and
molecular components is determined by magnetic pressure rather than gas pres-
sure. The motivation for this model is the unusually strong magnetic field mea-
sured in M17 through the Zeeman effect on the HI 21 cm line (Brogan et al. 1999).
Similar results were found for the neutral gas in front of the Orion nebula (the
so-called veil) by Abel et al. (2004). The idea is that soon after the formation of
a star cluster, starlight momentum and gas pressure from the ionized region com-
press the surrounding neutral gas, where the magnetic pressure builds up until it
can withstand the external pressure. See also Ferland (this volume).
A consequence of Pellegrini's model would be that the cosmic-ray flux in the
gas surrounding the star cluster is also enhanced, if the cosmic rays are trapped
in the magnetic field. The efficiency of this trapping depends on the geometry of
the magnetic field, in particular whether the field lines are open or closed. Direct
tests of this model are not easy because the Zeeman effect is difficult to measure.
However, an indirect test would be to look for correlations between the cosmic-ray
ionization rates of star-forming regions with their stellar luminosity.
Figure 1 shows recent observational estimates of the cosmic-ray ionization rate
in dense molecular clouds plotted against their luminosity. Although the estimates
are probably only accurate to order of magnitude, a trend does appear to be visible.
The possible correlation cannot just be a distance effect because the ionization
rates are local quantities derived from the ratio of two molecular abundances.
Thus, star formation may indeed influence the local cosmic-ray flux.
Besides variations with luminosity, the cosmic-ray ionization rate is also known
to differ between diffuse and dense clouds.
Ionization rates in diffuse gas are
factors of 3 -- 10 higher than those in dense clouds in the same region which are
exposed to the same incident cosmic-ray flux. This effect is shown by the open
symbols in Figure 1, with data from Le Petit et al. (2004) for ζ Per, McCall et al.
(2002) for Cyg OB2, Oka et al. (2005) for Sgr A and Geballe et al. (2006) for
IRAS 08572. The offset between the diffuse and the dense clouds must be due
to propagation effects, and scattering of cosmic rays off plasma waves may play
a role (Padoan & Scalo 2005), but only at low column densities (AV
∼ 10). In the
bulk of the clouds, the lifetime τ of the cosmic rays is probably limited by energy
losses, which scale as τ ∼ 2 × 105(nH /300cm−3)−1 yr (Gabici et al. 2007).
<
The authors thank Gary Ferland, Frank Israel and Sera Markoff for useful discussions.
References
Abel, N. P., Brogan, C. L., Ferland, G. J., et al. 2004, ApJ, 609, 247
Brogan, C. L., Troland, T. H., Roberts, D. A., & Crutcher, R. M. 1999, ApJ, 515, 304
4
Far-infrared observations of the interstellar medium
Fig. 1. Relation between cosmic-ray ionization rate and bolometric luminos-
ity.
Ionization rates are from Caselli et al. (2002), Van der Tak & van Dishoeck
(2000), Van der Tak et al. (2006) and Van der Tak et al. (2008); luminosities are from
Van der Tak et al. (2000), Lis & Carlstrom (1994) and Spaans & Meijerink (2007). The
luminosity of the pre-stellar core L1544 is an upper limit.
Caselli, P., Walmsley, C. M., Zucconi, A., et al. 2002, ApJ, 565, 344
Downes, D. & Eckart, A. 2007, A&A, 468, L57
Forster Schreiber, N., Genzel, R., Lutz, D., & Sternberg, A. 2003, ApJ, 599, 193
Gabici, S., Aharonian, F. A., & Blasi, P. 2007, Ap. Sp. Sc. , 309, 365
Geballe, T. R., Goto, M., Usuda, T., Oka, T., & McCall, B. J. 2006, ApJ, 644, 907
Grimm, H.-J., Gilfanov, M., & Sunyaev, R. 2003, MNRAS, 339, 793
Le Petit, F., Roueff, E., & Herbst, E. 2004, A&A, 417, 993
Lis, D. C. & Carlstrom, J. E. 1994, ApJ, 424, 189
Mauersberger, R. & Henkel, C. 1991, A&A, 245, 457
McCall, B. J., Hinkle, K. H., Geballe, T. R., et al. 2002, ApJ, 567, 391
Meijerink, R., Spaans, M., & Israel, F. P. 2007, A&A, 461, 793
Moran, E. C. & Lehnert, M. D. 1997, ApJ, 478, 172
Oka, T., Geballe, T. R., Goto, M., Usuda, T., & McCall, B. J. 2005, ApJ, 632, 882
Floris van der Tak et al.: Extragalactic H3O+: Some Consequences
5
Padoan, P. & Scalo, J. 2005, ApJ, 624, L97
Pellegrini, E. W., Baldwin, J. A., Brogan, C. L., et al. 2007, ApJ, 658, 1119
Seaquist, E. R., Lee, S. W., & Moriarty-Schieven, G. H. 2006, ApJ, 638, 148
Spaans, M. & Meijerink, R. 2007, ApJ, 664, L23
Van der Tak, F. F. S., Aalto, S., & Meijerink, R. 2008, A&A, 477, L5
Van der Tak, F. F. S., Belloche, A., Schilke, P., et al. 2006, A&A, 454, L99
Van der Tak, F. F. S. & van Dishoeck, E. F. 2000, A&A, 358, L79
Van der Tak, F. F. S., van Dishoeck, E. F., Evans, II, N. J., & Blake, G. A. 2000, ApJ, 537, 283
Walter, F., Weiss, A., & Scoville, N. 2002, ApJ, 580, L21
Zezas, A. 2006, Advances in Space Research, 38, 2946
|
astro-ph/0512359 | 1 | 0512 | 2005-12-14T10:15:42 | Micro-arcsecond light bending by Jupiter | [
"astro-ph"
] | The detectors designed for Gaia, the next ESA space astrometry mission to be launched in 2011, will allow to observe repeatedly stars very close to Jupiter's limb. This will open a unique opportunity to test General Relativity by performing many Eddington-like experiments through the comparison between the pattern of a starfield observed with or without Jupiter. We have derived the main formulas relevant for the monopole and quadrupole light deflection by an oblate planet and developed a simulator to investigate the processing of the Gaia astrometric observation in the vicinity of the planet. The results show that such an experiment carried out with the Gaia data will provide a new fully independent determination of the PPN parameter gamma by means of differential astrometric measurements and, more importantly, for the first time will evidence the bending effect due to the quadrupole moment with a 3-sigma confidence level. Given the accuracy of the experiment for the monopole deflection, this will permit to test alternative modelling of the light bending by moving masses. | astro-ph | astro-ph | Micro-arcsecond light bending by Jupiter
M T Crosta††and F Mignard†
†Observatoire de la Cote d’Azur, UMR CNRS 6202, Le Mont Gros, BP 4229, 06304
Nice cedex 4, France
E-mail: [email protected] and [email protected]
Abstract. The detectors designed for Gaia, the next ESA space astrometry mission
to be launched in 2011, will allow to observe repeatedly stars very close to Jupiter’s
limb. This will open a unique opportunity to test General Relativity by performing
many Eddington-like experiments through the comparison between the pattern of a
starfield observed with or without Jupiter. We have derived the main formulas relevant
for the monopole and quadrupole light deflection by an oblate planet and developed
a simulator to investigate the processing of the Gaia astrometric observation in the
vicinity of the planet. The results show that such an experiment carried out with
the Gaia data will provide a new fully independent determination of γ by means of
differential astrometric measurements and, more importantly, for the first time will
evidence the bending effect due to the quadrupole moment with a 3σ confidence level.
Given the accuracy of the experiment for the monopole deflection, this will permit to
test alternative modelling of the light bending by moving masses.
Submitted to: Class. Quantum Grav.
5
0
0
2
c
e
D
4
1
1
v
9
5
3
2
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
† Present address: Observatory of Turin, Strada Osservatorio 20, Pino Torinese (To), Italy
Micro-arcsecond light bending
1. Introduction
2
Several key questions of modern astrophysics regarding the formation and evolution of
the Milky Way will be clarified with the next space astrometry mission Gaia, approved
in 2000 as a cornerstone within the European Space Agency science program [1]. The
mission is now funded and will enter in the C/D phase in 2006 for a launch scheduled in
late-2011 and nominal operations over the next five years. Thanks to its high precision
(10 µarcsecond in angular measurements) and multi-epoch astrometry, Gaia will be able
to detect the relative positional change of a star resulting for the tiny curvature of the
light ray brought about by the gravitational pull of the Sun, and to a lesser extent by
the giant planets.
As far as the light deflection produced by the solar mass is concerned, the science
case has included from the outset an investigation of the value of the PPN parameter γ,
considered as a global unknown in the astrometric model. This parameter, equal to one
in General Relativity (GR), indicates the amount of space-time curvature produced by
a unit rest-mass and it is a fundamental part of the so called parameterized Post-
Newtonian (PPN) formalism [2], originally developed by Eddington for the famous
experiment during the solar eclipse in 1919. The PPN method aims to quantify the
violations of the Equivalence Principle by parameterizing deviations which modify both
local physical laws and large-scale gravitational phenomena, including the constancy of
the constants. PPN formalism is valid for a broad class of metric theories and includes
GR as a reference case, i.e. as a current standard theory of gravitation. The slow
motion and weak field limit allows to use PPN metric expansion as a function of several
parameters, varying from theory to theory, useful to discern gravitational experiments
in the Solar System. The PPN parameter γ is the most important, since all the other
parameters to all relativistic order within the alternative theories to gravity, converge
to their GR value in proportion to 1 − γ [3].
The most general formulation of these theories contains an arbitrary function of a
scalar field coupled to the stress-energy tensor in order to merge quantum mechanics with
gravity. Recent works on scalar-tensorial theories consistent with various cosmological
scenarios suggest the search for discrepancies from unity of γ at the levels of 10−5 to
10−7; more precisely, Damour and Nordtvedt [4] assume that there exist a cosmological
attractor mechanism towards GR quantified from deviation of γ from unity within the
above level of accuracy depending on the total mass density of the Universe. Then, an
experiment purposely designed could reveal the coexistence of the two mentioned fields
throughout the cosmological evolution up to the present where pure tensor gravity
explains the gravitational interaction from small to large scale.
In addition, some
dynamical models predict scalar fields dominating the current energy density of the
universe, which should contribute to the value of γ at the level of 10−7 to 10−9, thus
providing a universe evolution without the need of dark matter [3]. All this issues make
precision tests of gravity in space a fascinating challenge for the next decades, as in
the past the Eddington’s observations in 1919 of star line-of-sight did, confirming the
Micro-arcsecond light bending
3
amount of 1.′′72 as deflection angle predicted by GR. Nowadays, all present experimental
tests are compatible with the predictions of GR. In particular, experiments conducted
in Solar System have tested all the weak-field predictions of GR theory at better than
the 10−3 level and down to 2 × 10−5 level for γ as seen from the additional Doppler shift
in radio-wave beams connecting the Earth to the Cassini spacecraft when they passed
near the Sun [5].
New space projects will go deeper in experimental gravitation. The NASA Gravity
Probe B mission [6], whose first results are expected very soon, should directly measure
γ to better than the 10−5 level, although this is not the main objective of the experiment
and remains subjected to efficient instrumental calibrations. An early analysis of the
Gaia’s capabilities, indicates that its measurements should provide a precision of 5×10−7
for this parameter [7, 8], an improvement of more than two orders of magnitude to the
current best estimate mentioned above. This is quite comparable to the expectations
of the LATOR project [9], which is designed to use two optical interferometers between
two micro-spacecrafts and aims to determine 1 − γ at the level of 10−8.
Besides the determination of γ based on the global astrometry capabilities of Gaia,
the mission will also allow to carry out dedicated small field experiments from the
observations performed close to the surface of a giant planet like Jupiter or Saturn.
The light bending due to their gravitational field will be detectable in the visible and
we show in this paper that even the quadrupole field of Jupiter will be evidenced in
such an experiment.
It is therefore possible to carry out the same kind of eclipse
experiment made by Dyson, Eddington, and Davidson in 1920, but now by comparing
stellar positions in the immediate vicinity of the planets. Not only will this experiment
be a further confirmation of a GR prediction, but it will also help understand the very
difficult question of the light-bending by a moving source and that of the propagation
of gravitation.
Using Jupiter to test GR, Kopeikin and Fomalont claimed they have determined
the speed of gravity trough the VLBI ‡ astrometric measurement of Shapiro time delay
of light from the Quasar J0842+1835, as its image passed within 3.7 arcmin of the planet
(about 10 Jupiter radii, a far cry from what Gaia can achieve) [10]. According to the
post-Minkowskian formalism assumed [11], the gravitational field due to a moving body
propagates with finite speed; consequently, it influences a photon with some retardation.
The speed of gravity enters as an extra tiny velocity-dependent corrections to the Shapiro
time delay formula of the order of 4.8 ps. At present there is no general consensus on
the results, especially upon the method to extend GR to the case where the velocity
of gravity differs from the velocity of light [12, 13]. In particular, given the observed
limits on γ, the Fomalont-Kopeikin measurements are not accurate enough yet to
determine the speed of gravity [12]. Some authors interpret the result within the Lorentz
transformation properties of the weak gravitational field and consider it as a violation
not only of the Lorentz invariance but also of the Galilean one [14,15]. Paper [16] settles
‡ VLBI is the acronym for Very Long Baseline Interferometry
Micro-arcsecond light bending
4
δΦpN
1.′′75
µas
83
493
574
26
116
16280
5772
2081
2535
7
δΦQ
δΦmax
∼ 1 µas
(180◦)
µas
–
–
0.6
–
0.2
239
94
25
9
–
9′
4.5◦
178◦
9◦
25′
90◦/3′
18◦/51′′
72′/6′′
51′/3′′
8′′
Sun
Mercury
Venus
Earth
Moon
Mars
Jupiter
Saturn
Uranus
Neptune
Pluto
Light deflection amounts at 1 µarcsec due to the planets for a photon
Table 1.
crossing the Solar System: pN is for post-Newtonian order and Q for the quadrupolar
moment. The values are computed for grazing rays in the Gaia observing geometry;
figures in the last column give the angular distances between the perturbing body and
the star at which the effect is still 1 µarcsecond, with Gaia at the Sun-Earth L2. Where
two values are reported they refer, respectively, to pN and quadrupole effect.
the significance of the Jupiter experiment as a simply standard aberration of light which
propagates across an active medium with an effective index of refraction induced by the
gravitational field of a lens in motion: the tiny corrections to the Shapiro time delay
should be included into the fitting of microlensing events. However, even if there is
still a controversy about the question of whether the results depend on the speed of
gravity [17] or on the abberation of light [14, 16, 18], there is a general agreement that
this measurement is a remarkable precise VLBI detection of a post-Newtonian effect
due to a planet.
The possibility to apply methods of relativistic astrometry to test GR has
prompted several groups to model highly accurate angular observations in the relativistic
framework. New formalisms have been proposed ( [11, 19–21] and reference therein) to
tackle the relativistic problem of the light path reconstruction, which, at least at the
µas accuracy, implies taking into account all the contributions not only due to the
bulk mass of the Solar System bodies, but also to their quadrupole moment and to the
time-dependent gravitational field generated by their motions (see table 1).
In particular, two papers [22] and [23] have gone through the post-Newtonian
treatment of the light propagation in the case of an isolated axisymmetric body, in
order to take into account the influence of the multipole moments. The former is based
Micro-arcsecond light bending
5
on a mathematical solution of geodesic equation whereby mathematical expressions for
the light deflection in any order of multipole perturbations are obtained. In the second
paper, this contribution stems from the prior computation of the time transfer function
between two point located at a finite distance. However neither approach has provided
hints on how to apply the intricate formulas to a practical case like the one faced in the
Gaia modelling.
Having in mind the capabilities of Gaia to observe stellar sources very close to
Jupiter’s limb (at a fraction of the planet radius), we have evaluated the light deflection
produced by an oblate planet on grazing photons coming from distant stars. This study
is part of a wider project called GAia Relativistic EXperiment (GAREX), which aims
to investigate all possibilities to test GR with Gaia measurements. In the solar system,
this mission will mainly carry out:
• light deflection experiments, divided into (i) global astrometry, in particular highly
accurate determinations of γ by observing the change in star position at different
angular distance to the Sun; (ii) small field experiments, with the examination of
the light propagation by means of differential measurements of stellar positions near
the planets;
• perihelion precession effect, related to the determinations of PPN parameter β §
from the orbit fitting of several thousands minor bodies [24].
In this paper we concentrate on the case of the light bending by Jupiter at the
microarcsecond level and its detection by the small field astrometric observations with
Gaia. In [25] we have already explored the favorable circumstances for Gaia in five years
of continuous observations, approximately from 2012 to 2017, to detect the quadrupole
light bending effect predicted by GR, but never observed. For Jupiter the magnitude of
the monopole deflection for a grazing ray is ∼ 16 milliarcsecond, to which a component
from the quadrupole moment is superimposed with an amplitude of ∼ 240 micro-
arcsecond, as showed in table 1. This secondary deflection has a very specific pattern as
a function of (i) the position of the star with respect to the oblate deflector and (ii) the
orientation of its spin axis. In section 2 we derive the relevant formulas to express this
light bending effect using the PPN formalism at large and small angle from the planet.
In section 3 we describe all the steps needed to simulate the small field experiment when
observing stellar background in vicinity of Jupiter. Finally, in section 4, we discuss the
data processing of a set of Montecarlo runs and the results in the determination of the
PPN Parameter γ and the quadrupole effect from Jupiter.
2. The light deflection produced by an axisymmetric planet
Most stellar sources, planets and observers have a small velocity compared to the
velocity of the light and a weak inter-body gravitational field exists inside bound
systems. Moreover, if we consider isolated distribution of matter, the geometry can
§ The PPN parameter β measures the non-linearity in the superposition of gravitational fields.
Micro-arcsecond light bending
6
be assumed Minkowskian, asymptotically flat far away [26]. Let us consider, then, a
set of PPN coordinates {xα} (α = 0, 1, 2, 3) and a solar system locally perturbed by
isolated stationary axisymmetric masses. To the post-Newtonian (pN) accuracy, the
spatial part of the geodesic equation of a light ray coming form a distant star can be
easily transformed into [2]:
d2xi
dx02 = U,i 1 + γ δjk
dxj
dx0
dxk
dx0! − 2(1 + γ)
dxi
dx0 δjk
dxj
dx0 U,k! ,
(1)
where U = M/q(δijxixj) (i, j, k = 1, 2, 3) is the Newtonian potential (in geometrized
units) and U,i its partial derivative with respect to the spatial coordinates. A solution
of the above equation can be get by tracing a Newtonian zero order straight-line xi
N
plus all the relativistic deviations xi
D, i.e.
xi = xi
N + xi
D.
Then, by decomposing xi
to xi
N , equation (1) becomes
D into two components, parallel (xi
Dk
(2)
) and perpendicular (xi
D⊥)
D⊥
d2xi
dx02 = (1 + γ)[U,i −ti(δjktjU,k )] = (1 + γ)∇⊥U,
(3)
where t = dxi/dl is the unit tangent vector on the unperturbed light path in the direction
of the observer. Here the symbol ∇⊥ indicates the derivative perpendicular to the light
ray, assumed to pass outside the matter distribution. Placing the origin of coordinates
xi
0 at the center of an axisymmetric planet, i.e. z considered as its axis of maximum
moment of inertia, located between the star and the observer, the positional vector of
the photon with respect to the principal axes centered at the planet can be expressed
as:
ri = xi − xi
0 = tiℓ − nib,
(4)
where ℓ represents the length of the light path and n the radial direction perpendicular to
the unperturbed ray pointing towards the center of gravity, along the impact parameter
b.
The light deflection is given by the deviation of the tangent vector t from the
straight-line along the photon’s path as,
which leads with (3) to,
dℓ
dℓ
∆Φ ≡ Z dt
∆Φ = (1 + γ)Z ∇⊥U dl.
r −
∇⊥U = "−
+(cid:18)−
+(cid:18)−
3M
3M
b
By taking the gravitational potential up to the quadrupole term one gets,
!
2
3M
r4 J2R2 5 cos2 θ − 1
M
r2 +
r4 J2R2 cos θ(cid:19) (z · n)(cid:21) n
r4 J2R2 cos θ(cid:19) (z · m)m
(5)
(6)
(7)
Micro-arcsecond light bending
7
where r and θ are respectively the radial distance and the co-latitude (measured from
the planet north pole) of a field point on the light trajectory, J2 is the dimensionless
coefficient of the second zonal harmonic, R the radius and M the mass of the planet,
and, finally, m represents the orthoradial component (see figure 1 and figure 2).
z
r
P
n
b
a
t
m
u
O
Figure 1. Geometry of light deflection due to a planet (P): the spin axis of the planet
z is out of the plane; t represents the unit tangent vector from a distant star (S) to
the observer (O) on the unperturbed light trajectory; u is the unit direction from O
to P along their distance a; finally, χ is the angle S OP, and b the impact parameter.
m
-n
cos(d )
Figure 2. Light deflection by a planet: tangent plane on the sky. The position of the
star is displaced both in the radial (-n) and orthoradial m directions. The spin axis of
the planet z (not shown here) does not lie in this plane in general.
The length ℓ of the photon path can be scaled by the impact parameter into a
dimensionless parameter as follows
dℓ = bdλ,
(8)
c
q
D
a
D
d
Micro-arcsecond light bending
so the radial distance becomes
r = b(1 + λ2)1/2.
8
(9)
Each integral entering (6) must be computed with λ running positively in the same
direction as the photon from λ = −∞ to λ = 1/ tan χ, with χ standing for the angular
separation between the directions star/observer and observer/planet (figure 1). At the
closest approach on the unperturbed ray one has λ = 0. The explicit expressions of the
integrals are given in appendix A. After some algebra, the light deflection vector is split
into two components, the first one along n and the second one along m, both including
the monopole and the quadrupole contribution of the planet in function of the angular
separation χ:
∆Φ = ∆Φ1n + ∆Φ2m,
where, precisely,
∆Φ1 = (1 + γ)
2M
b ((1 + cos χ) + J2
−2(1 + cos χ +
1
2
cos χ sin2 χ +
+(sin3 χ − 3 sin5 χ)(n · z)(t · z)
3
4
R2
b2 (cid:20)(1 + cos χ +
1
2
cos χ sin2 χ)
cos χ sin4 χ)(n · z)2
− (1 + cos χ +
1
2
cos χ sin2 χ −
3
2
cos χ sin4 χ)(t · z)2(cid:21)(cid:27)
and
∆Φ2 =
(1 + γ)MJ2R2
b3
(cid:20)2(1 + cos χ +
1
2
cos χ sin2 χ)(n · z)(m · z)
(10)
(11)
(12)
+ sin3 χ(m · z)(t · z)i .
The first term in the radial component (that along n) is the classical monopole deflection,
widely used in astronomy. All the other terms factored by J2 come from the quadrupole
of the planet. One can specialize these expressions to near-grazing rays as, unlike with
the solar deflection, the effect is too small to be observed at large angle from the planet.
Hence when χ ≪ 1 this leads to the more convenient and accurate enough formulas,
∆Φ1 =
∆Φ2 =
2(1 + γ)M
b
"1 + J2
R2
b2 (cid:16)1 − 2(n · z)2 − (t · z)2(cid:17)#
4(1 + γ)MJ2R2
b3
(m · z)(n · z).
(13)
(14)
which are similar to a derivation obtained in [27] in the case of the Sun.
The deflection vector depends on the orientation of the spin axis of the planet and,
moreover, on the direction of the star with respect to the planet as seen in the terms
proportional to (n · z) and (t · z). In general the quadrupole deflection vector depends
both on the impact parameter and on the direction of the light source with respect to the
spin axis projected on the plane of the sky. However, for a given impact parameter, its
modulus (10) depends only on the star’s direction with respect to the planet’s reference
Micro-arcsecond light bending
9
frame. This property is established in Appendix B and has been used thoroughly to as
an additional check to the computer implementation of the simulation (see section 3.2).
One should note that the above formulation has left aside and neglected refinements in
the modelling like the motion of Jupiter, the retarded effect on the light propagation
and mulipolar terms of higher orders in the potential expansion.
Considering γ = 1 is equivalent to saying that equations (13)-(14) constitute the
reference modelling in the standard GR. Therefore, in the same way as one introduces
γ in the monopole deflection to characterize departures from GR, we multiply J2 by a
dimensionless parameter ǫ whose value is 0 if no effect is seen (say this may correspond to
a lack of coupling between gravity and electromagnetism seen in the multipole moments)
and 1 for the standard GR. Hereafter we call this parameter the Quadrupole Efficiency
Factor (QEF) and the goal is to see whether it is significantly determined from the Gaia
observations. Although ǫ is not a PPN parameters, it represents the first level of a
post-Newtonian test whose goal is to determine whether the effect is detected,i.e. if one
can say that ǫ is equal to unity, with a certain uncertainty. Thus, for the processing of
the experiment of light deflection near the planet, we adopt the following formulation,
∆Φ1 =
∆Φ2 =
2(1 + γ)M
b
"1 + ǫJ2
R2
b2 (cid:16)1 − 2(n · z)2 − (t · z)2(cid:17)#
4M(1 + γ)ǫJ2R2γ
b3
(m · z)(n · z).
(15)
(16)
including the two unknowns γ and ǫ, which are equal to 1 in the simulation.
3. Small field experiments with Jupiter
The one GAREX experiment we consider in this paper is a light deflection in the vicinity
of Jupiter. The goal is to measure in a fully independent way the two parameters γ
and ǫ expressed in light deflection formulae (15) and (16), together with their associated
errors, knowing the astrometric accuracy that will be achieved by Gaia. As J2 is well
known from space probe tracking with a precision better than 10−5, we will here consider
it as a constant, and infer instead an estimate of the QEF factor ǫ.
3.1. Ephemerides
First, we have investigated what are the favorable circumstances to perform experiments
using Jupiter during the mission lifetime [25]. This depends on the number of times
Jupiter will cross one of the astrometric fields during the mission and on the stellar
density in its immediate surroundings during these observations. At present, the details
of the former can only be known statistically, as the precise initial conditions of the sky
scanning law remain unknown. On the other hand the sky background can be assessed
with more certainty as it depends only on the motion of Jupiter between 2012 and
2018, and not very accurately on when the observations take place, provided they are
reasonably scattered during the observing period. The ephemerides of Jupiter have been
Micro-arcsecond light bending
10
computed in galactic coordinates between these two epochs for an observer orbiting the
Sun-Earth L2 in the same way as Gaia will do. The observability conditions are shown
in figure 3 with the visibility periods when the angular distance to the Sun makes the
observations feasible with Gaia (no observations can be made around conjunctions and
oppositions). Galactic coordinates are used just because the primary dependence of the
stellar density is on the galactic latitude.
Figure 3. Galactic latitude (b) of Jupiter versus the observing time (in years). The
solid line represents the galactic latitude of Jupiter (as seen from Gaia) over the years
2011-2018 and the highlighted patches correspond to the visibility periods when the
angular distance to the Sun makes the observation possible.
We have chosen the date 2012 as the beginning of the simulations, since Gaia will
take at least six months to be operational at the Lagrangian point L2 of the Earth-
Moon-Sun system. Jupiter will cross the galactic plane in mid-2013 and, consequently,
we expect favorable observations of the planet in front of a dense stellar background as
shown in figure 4. The plot gives the average number of stars per square degree around
Jupiter from 2012 to 2018 as seen from Gaia.
3.2. Simulation of the experiment
The principle of the simulated experiment is to generate a series of pairs of observations
over the mission length of 5 years, so that in each pair Jupiter is, or is not, at the center
of the stellar field. In both cases astrometric observations similar to that expected from
Gaia are computer generated, with the proper noise function of the star brightness. The
number of stars simulated in each field is determined by the galactic latitude of Jupiter
at the observation time. Then the two star patterns are compared and the differences in
along-scan positions (the only direction of accurate measurements with Gaia) are fitted
Micro-arcsecond light bending
11
Figure 4. Mean number of stars per square degree down to the magnitude limit
V=20 in the field of Jupiter during the Gaia operations. The dashed lines represent
the period when Jupiter crosses the galactic plane and the central one the epoch when
the maximum number of stars is observed.
to the model with the two unknowns γ and ǫ. This is an idealized version of the Gaia
procedure as the two fields (with or without Jupiter) will not be observed exactly a the
same time, leading to correction for the relative orientations of the fields and for the
proper motion of the stars. But these are technical details that will be carefully handled
during the data processing and have no impact on the feasibility study demonstrated in
this paper. The numerical parameters used in the simulation are listed in table 2.
In order to use equations (15)-(16) in the simulation, we have fixed the origin
of the coordinate system at Jupiter’s center and assumed that it coincides with the
center of the astrometric focal plane. Given the galactic coordinates (l, b) of Jupiter we
extracted the relevant stellar density in the magnitude range V = 12.0 − 20.0 from a
realistic galactic model based in actual star counts fitted to convenient mathematical
expressions as a function of the galactic latitude and longitude. The magnitude range
is determined by the rejection of saturated stars (brighter than V=12.0) on the bright
side and by keeping stars observable with Gaia with a reasonable astrometric accuracy
on the faint side. One must keep in mind that the number of close approaches between
Jupiter and stars brighter than V ≃ 12 will be a rare event, albeit very important in
this context, but not liable to statistical treatment. It must be handled on a case by
case basis with a real sky and this will be done later when the final parameters of the
scanning law are known. After having computed the number of stars per interval of
0.5 magnitude inside the field of view at each observational epoch, we generated the
observed stars in the field of view. For the fields with Jupiter we have computed the
Micro-arcsecond light bending
12
Table 2. Main parameters used for the simulation.
parameter
number of observations
area of the field-of-view
magnitude range
RJ
M = GMJ /c2
Jupiter J2
γ
ǫ
σpos along the scan at V = 15
σpos across the scan at V = 15
numerical value
90
0.6 deg × 0.6 deg
12.0 to 20.0
7.13 × 107 m
1.41 m
0.014736
1
1
100µas
300µas
proper stellar directions by taking into account the full deflection due to the monopole
and the quadrupole. Then for any pair and any epoch we have added a gaussian noise in
both coordinates. The standard deviation of the noise is determined by the magnitude
of the star and the coordinates (along- or across-scan) with [1]:
σals = a × 100.2(V −15)
(17)
for stars with magnitude V > 12.5. The coefficient a is the scaling factor giving the
single observation astrometric accuracy for a star of 15 magnitude. For stars brighter
than 12.5, σals is constant and chosen to be 30 µas for the measurements along the scan
and 100 µas across. We have used the property expressed in (B.3) to cross-check the
contribution of the quadrupole term in the simulation. Other checks were also performed
in different reference frames that helped improve the reliability of the simulator. Figure 5
and figure 6 illustrate several important features of the simulated data.
The plots in figure 5 show the absolute values of the monopole and of the quadrupole
deflections, obtained with the simulator for the different epochs and impact parameters.
Figure 6 is an example of the observed field corresponding to mid 2013 when Jupiter
crosses the Galactic plane. Even if this date represents a favorable observation, the
relatively limited number of stars close to Jupiter does not lead to an obvious signature
of the stellar displacements produced, in particular, by the quadrupole effect. For this
reason we have also plotted the cumulative effect in figure 7 by combining all the epochs.
As expected one sees clearly the radial nature of the monopole deflection and the more
complex pattern of the quadrupole effect.
Micro-arcsecond light bending
13
Figure 5. The monopole (left-hand side) and quadrupole (right-hand side) deflections
over all epochs of Jupiter’s visibility for all stars up the magnitude limit V=20, where
1 < b/RJ ≤ 2 (full square), 2 < b/RJ ≤ 3 (open square), 3 < b/RJ ≤ 4 (open
triangle), 4 < b/RJ ≤ 5 (full triangle), 5 < b/RJ ≤ 6 (cross), b/RJ > 6 (dots).
Figure 6. Observer view of the monopole (left-hand side) and quadrupole (right-hand
side) light deflection vector field around Jupiter (circle) in mid 2013. The scales of 10
marcsecond for the monopole and 20 µarcsecond for the quadrupole are shown on the
lower left of the plots, while the dotted line indicates the direction of the Jupiter’s spin
axis with respect to the line-of-sight.
Micro-arcsecond light bending
14
Figure 7. The monopole (left-hand side) and quadrupole (right-hand side) stellar
vector field around Jupiter (circle) over the epochs between 2012 and 2018 from the
observer point of view. The scales of 10 mas for the monopole and 20 µas for the
quadrupole are shown on the lower left of the plots.
4. Measurement of the deflection
4.1. Processing the observations for γ and J2
The observed quantities used to fit the relativistic effect in the light deflection are
the along-scan positional differences of the stars referred to a common origin, namely
the center of gravity of Jupiter. More precisely, we have estimated the measurements
(∆Φals) along scan between two transits: the first one with Jupiter (Φ+J ) and the
second one without Jupiter (Φ−J ), i.e.:
∆Φals = Φ+J − Φ−J .
(18)
As stated by (18), the method rests only on the comparison between small fields taken
at two distinct epochs (each point on the sky will be mapped at least three times during
six months) and it is largely independent from the attitude reconstruction of Gaia. This
makes possible, in estimating the final accuracy, to consider only random errors and not
the systematic attitude errors shared by all the sources at the two epochs.
The set of n observations can be considered as a system of n equations where the
unknowns are the parameters γ and ǫ. The condition matrix is calculated by evaluating
the partial derivatives with respect to γ and ǫ in (18). The large number of stellar
measurements, about 160,000 for 90 Jupiter observations scattered over the 5-year
mission is solved for the two unknowns with a weighted least squares procedure using
singular value decomposition. The off-diagonal terms of the final covariance matrix are
not significant, indicating a very weak correlation between the two fitted parameters.
Micro-arcsecond light bending
15
Finally anomalous observations are filtered out with an iterative Student ratio test on
the residuals.
The estimate of the accuracy with which Gaia could determine γ and ǫ from small
field astrometry has been estimated by running more than 100 times this numerical
simulation with random drawings stochastically independent in each experiment. The
average of the 100 values of γ and ǫ prove that there is no bias in the determination as
seen in figure 8. As for the precision, the scatter was taken as a more robust way to
assess the accuracy than the formal error of the diagonal terms of the covariance matrix.
The mean and scatter have been evaluated by using the usual point estimates valid for
nrun ≫ 1:
< γ > =
< ǫ > =
1
nrun
1
nrun
σγ
σǫ
= vuut
= vuut
nrun
1
1
nrun
γ(i)
ǫ(i)
nrun
nrun
nrun
Xi=1
Xi=1
Xi=1
Xi=1
nrun
γ(i)2 − < γ >2
ǫ(i)2 − < ǫ >2.
(19)
(20)
(21)
(22)
Figure 8. The distribution of the mean values of γ (left-hand side) and ǫ (right-hand
side) over 150 Montecarlo simulations. The bars are the standard deviations.
Micro-arcsecond light bending
4.2. Results
16
The main results are listed in table 3 with the average and scatter on γ and ǫ for three
independent Montecarlo runs.
Table 3. Results for the mean value and scatter of γ and ǫ in Montecarlo experiments
of different lengths.
< γ >
0.9910
1.0000
1.0000
σγ
< ǫ >
σǫ
runs
1.0295×10−3
1.1473×10−3
1.1456×10−3
1.012
0.992
0.986
0.335
0.358
0.361
50
100
150
It is found with all the assumptions described in the previous sections that γ can
be determined from the monopole light deflection by Jupiter with an uncertainty of
1.1 × 10−3. This is three times better than the optical determination achieved by
Hipparcos, the previous astrometric ESA mission, with stellar astrometry at large angles
from the Sun [28] and 5 million observations of stars. The interest rests not only on the
accuracy achieved (the global astrometry should reach four orders of magnitude better
with solar light-bending), but on the fact that (i) it can be done with a planet, (ii) it
is a prerequisite to the detection of the quadrupole effect in the residuals, (iii) it opens
the way for testing the accurate modelling of the deflection by a moving body (it is
stationary in our experiments).
The final result in ǫ confirms that the gravitational effect due to J2 is detectable
with Gaia (albeit marginally), with typically ǫ = 1 ± 0.35, that is to say a 3-σ detection.
The effect of J4 is negligible since it is much smaller than the random noise on most
stars. Using J4 ∼ −0.0006 ∼ 4J2/100 this results for a grazing ray into a deflection of
10µas. In addition the effect of J4 decreases as (RJ /r)5 instead of (RJ /r)3 and very few
stars close to the planet limb would contribute to the signal.
4.3. Best strategy for the actual experiment
The above results concern a statistical analysis over all the epochs of observations during
five years, it is however useful to see whether some controlled parameters are more
favorable than others to observe the light bending effect and if one can restrict the
risk of systematic errors. First of all we have considered the evolution of the errors
on γ and ǫ with the limiting magnitude, for various impact parameters and for all the
epochs. As we include fainter stars, they get more numerous but each with a smaller
weight due to the degradation of the astrometric accuracy.
It was not obvious what
effect would overcome the other: could many faint stars contribute significantly or not?
Micro-arcsecond light bending
17
The risk being to increase significantly the chance of systematic deviation for almost no
improvement in the solution.
The results are plotted in figure 9, where the precision is shown when only stars
brighter than V are used in the fit. Each curve corresponds to a limitation in the size of
the field of view around Jupiter. It’s clear (and not surprising) that using the full Gaia
field provides the best results, in particular for γ just because the monopole deflection
decreases more slowly with increasing impact parameter than the quadrupole’s. However
going further than 30RJ brings very marginal improvement and this can be used to
determine what a small field experiment means in practice. Conversely below 10 RJ
the number of stars decreases very quickly and at 5 RJ we usually have no star brighter
than V=17.5. As for the magnitude it seems that one could also reject stars (in case of
calibration problems for example) fainter than V = 16 without much loss as the final
precision on γ and ǫ does not improve very much when they are included. That’s a
very nice feature when it comes the time to decide on whether faint stars are useful or
not. The graphs show also a degradation in precision when only the brightest stars are
considered, probably because the number of stars becomes too small to be significant and
formula (17) inserts a constant error in this case. Then another interesting experiment
Figure 9. Evolution of the errors on γ and on ǫ with the magnitude and for various
impact parameters. Full circles indicate the complete field corresponding to 45 RJ ,
open circles 30 RJ , full squares 20 RJ , open squares 15 RJ , full triangle 10 RJ and,
open triangle 5 RJ .
has been run to test what would be the result if one selected only the few epochs during
which we have the maximum number of stars in the background field, that is to say
around 2013 when Jupiter is close to the Galactic plane. To this purpose, figure 10
shows again the evolution of the errors with magnitude, when all observations at every
Micro-arcsecond light bending
18
epoch are included (full circles) or only during the crossing of the Galactic plane in
2013-2014 (open circles with only 20 epochs instead of 90 in the nominal experiment).
Not surprisingly the errors are larger when considering only the best epochs, but not
dramatically larger, showing that a strategy may be applied to perform valuable and
dedicated experiments during the first two years of the mission, to validate the concept
and get already significant results.
Figure 10. The evolution of the errors with magnitude, keeping the whole Gaia field
for all epochs (full circles) and for 2013 (open circles), when we have the maximum
number of stars.
Finally one should also look at the passage of Jupiter in the middle of a concentrated
open cluster with stars brighter than V = 13. There are many such clusters on the
ecliptic in the direction of the Sagittarius, that Jupiter will cross in 2019. This is after
the nominal mission completion, but well within the possible extended mission permitted
by the consumable. This kind of experiment would obviously add in the science case
to support this extension of the operations. To this aim we have run the simulator
with bright stars and a surface density comparable to that found in these open clusters.
Results for the determination of γ and ǫ are given in table 4 and table 5 with all the stars
of 12 or 13 magnitude respectively. Two realistic concentrations have been considered
around Jupiter corresponding to fields of 3 and 6 arcmin around the planet. One sees
that a single experiment can do almost as well as the 5-year mission, provided the number
of bright stars in the cluster is large enough. Note that we have excluded a priori
all the stars with magnitude range between 14 and 16, which statistically contribute
significantly to the precision. At this point this is just an indication showing that a
detail investigation on specific clusters is worth doing.
Micro-arcsecond light bending
19
Table 4. Precisions computed after 150 Montecarlo runs selecting special fields in
2013 and increasing artificially the number of stars (n*) (V=12).
b = 7RJ
b = 15RJ
n*
σγ
σǫ
1
5
13
20
27
34
7
23
39
63
79
119
7.41 ×10−3
3.37 ×10−3
1.62 ×10−3
1.39 ×10−3
1.06 ×10−3
9.31 ×10−4
14.54
2.73
0.69
0.34
0.32
0.27
4.48 ×10−3
2.12 ×10−3
1.70×10−3
1.31 ×10−3
1.26 ×10−3
9.16 ×10−4
9.09
1.51
0.94
0.62
0.51
0.30
Table 5. Precisions computed after 150 Montecarlo runs selecting special fields in
2013 and increasing artificially the number of stars (n*) (V=13).
n*
9
29
49
66
91
b = 7RJ
b = 15RJ
2.49 ×10−3
1.16 ×10−3
9.53 ×10−4
7.50 ×10−4
6.50 ×10−4
44
67
81
97
127
1.88 ×10−3
1.56 ×10−3
1.43 ×10−3
1.17 ×10−3
1.09 ×10−3
σγ
σǫ
1.51
0.37
0.26
0.20
0.15
1.28
0.62
0.58
0.41
0.35
Micro-arcsecond light bending
5. Conclusion
20
The objectives of this paper has been to study a new way to determine the PPN
parameter γ by exploiting a method of differential positional measurements around
Jupiter and to assess the detection of the light bending from the quadrupole moment of
Jupiter. Thanks to the high astrometric accuracy to be achieved by the ESA astrometry
mission Gaia and the repeated observations over five years this will be feasible as the new
design of Gaia’s optical instrument allows to process stellar observations very close to
the surface of giant bodies. We have shown that the deviation to GR with the monopole
deflection can be assessed to 10−3 with Jupiter only, a result in the visible better than
the optical accuracy already achieved by Hipparcos with the stars and the solar light
bending. With the same observations, the quadrupole light deflection will be detectable
for the first time with a 3-σ confidence level.
Although we have designed an ideal and simplified experiment it includes realistic
observations of Jupiter feasible with Gaia, taking into account the astrometric accuracy
with the star’s magnitude deduced from the current error budget analysis. Whereas the
Gaia concept and its design rest entirely on the global astrometry we have given for the
first time a realistic figure on the true strength of Gaia to carry out also relativity testing
with small field astrometry and propose some strategies to carry out this experiment in
the best conditions. When the initial conditions of the Gaia scanning law are known,
the same principles will be applied using a real stellar distribution. The method will be
also extended to the observations of Saturn, at least for the monopole effect.
Acknowledgments
MTC acknowledges financial support from the Henri Poincar´e Fellowship during her
stay at the Observatoire de la Cote d’Azur (Cassiop´ee Department). We wish to thank
the members of the Relativistic and Reference Frame Working Group of Gaia for their
valuable comments and collaboration.
Appendix A
This appendix shows
computed to obtain
expressions (11) and (12) for the deflection vector ∆Φ after having scaled the photon
length by the impact parameter:
fundamental
integrals
the
set of
dλ
(1 + λ2)3/2 =
λ
(1 + λ2)1/2
dλ
(1 + λ2)5/2 =
λ/3
(1 + λ2)3/2 +
2λ/3
(1 + λ2)1/2
dλ
(1 + λ2)7/2 =
λ/5
(1 + λ2)5/2 +
4λ/15
(1 + λ2)3/2 +
8λ/15
(1 + λ2)1/2
λdλ
(1 + λ2)5/2 = −
1/3
(1 + λ2)3/2
Z
Z
Z
Z
Micro-arcsecond light bending
21
λdλ
(1 + λ2)7/2 = −
λ2dλ
(1 + λ2)7/2 = −
Z
Z
1/5
(1 + λ2)5/2
λ/5
(1 + λ2)5/2 +
λ/15
(1 + λ2)3/2 +
2λ/15
(1 + λ2)1/2
Appendix B
Let us consider the angle ϕ, ϑ of the Jupiter axis with respect the orthonormal directions
n, and t, respectively. From the left hand-side of equation (13) we get
1 − 2(n · z)2 − 2(t · z)2 = − sin2 ϑ · cos 2ϕ,
and from that one of equation (14)
2(n · z)(m · z) = sin2 ϑ sin 2ϕ.
(B.1)
(B.2)
Then, the modulo of the quadrupole deflection term, let’s say ∆Φquad, in equation (10))
will depend only on sin2 ϑ, namely on the angle between the z axis of Jupiter and the
direction from the star to the observer
∆Φquad =
2(1 + γ)MJ2R3
b3
sin2 ϑ
(B.3)
Clearly, when the z lies on the {n, m} plane, i.e. z perpendicular to the line-of-sight,
the quadrupole term will give its maximum contribution.
For a star in direction α with respect to the orthonormal axis {m, n, t} the modulo
has a fixed value depending only on the star’s direction of arriving with respect to the
above frame centered on Jupiter, namely:
∆Φquadn = ∆Φquad cos 2α
∆Φquadm = ∆Φquad sin 2α.
(B.4)
(B.5)
References
[1] 2005 Proc. of the Symposium The Three-Dimensional Universe with Gaia (Paris) ESA-SP-576
[2] Will C M 1993 Theory and experiment in gravitational physics (Cambridge University Press)
[3] Turyshev S G, Williams J G, Nordtvedt K Jr, Shao M and Murphy T W 2003 Lect. Notes Phys.
648 301-320
[4] Damour T and Nordtvedt K 1993 Phys. Rev. Lett. 70 2217
[5] Bertotti B, Iess L and Tortora P 2003 Nature 425 374
[6] Gravity Probe B Project at Stanford University, http://einstein.stanford.edu/
[7] Mignard F 2002 Gaia: A European Space Project vol. 2 (EAS Publications Series) p 105
[8] Vecchiato A, Lattanzi M G, Bucciarelli B, Crosta M T, de Felice F and Gai M 2003 A&A 399 337
[9] Turyshev S G, Shao M and Nordtvedt K L 2004 J.Mod.Phys. D13 2035-2064
[10] Fomalont E B and Kopeikin S M 2003 Astrophys.J. 598 704-711
[11] Kopeikin S M and Mashhoon B 2002 Phys. Rev. D 65 064025
[12] Carlip S 2004 Class. Quantum Grav. 21 3803-3812
[13] Will C M 2003 Astrophys.J. 590 683-690
[14] Asada H 2002 Astrophys.J.Lett. 574 L69
[15] Samuel S 2004 Int. J. Mod. Phys.D 13 1753
Micro-arcsecond light bending
22
[16] Frittelli S 2003 MNRAS 344 L85-L87
[17] Kopeikin S M 2004 Class. Quantum Grav. 21 3251-3286
[18] Pascual-S`anchez J F 2004 Int.J.Mod.Phys. D 13 2345
[19] Klioner S 2003 Astron.J. 125 1580
[20] de Felice F, Crosta M T, Vecchiato A, Lattanzi M G and Buciarelli B 2004 Astrophys.J. 607 580
[21] Le Poncin-Lafitte C, Linet B and Teyssandier P 2004 Class. Quantum Grav. 21 4463
[22] Kopeikin S M 1997 J.Math.Phys. 38 2587
[23] Le Poncin-Lafitte C and Teyssandier P 2004 Proc. of the Journ´ees Syst`emes de R´ef´erence Spatio-
Temporels (Obs. Paris) p 204-209
[24] Hestroffer D and Berthier J 2005 Proc. of the Symposium The Three-Dimensional Universe with
Gaia (Paris) ESA-SP-576 p 297-300
[25] Crosta M T and Mignard F 2005 Proc. of the Symposium The Three-Dimensional Universe with
Gaia (Paris) ESA-SP-576 p 281-284
[26] Misner C W, Thorne K S and Wheeler J A 1973 Gravitation (Freeman)
[27] Epstein R and Shapiro I 1980 Phys. Rev. D 22 12
[28] Froeschl´e M, Mignard F and Arenou F 1997 Proc. of the ESA Symposium Hipparcos - Venice 97
ESA-SP-402 p 49-52
|
0709.3730 | 1 | 0709 | 2007-09-24T10:22:24 | Interaction and ablation of fall-back disks in isolated neutron stars | [
"astro-ph"
] | An analysis of ablation processes is made for a fall-back disk with inner and outer radii external to the neutron-star light cylinder. The calculated ablation rate leads, with certain other assumptions, to a simple expression relating the inner radius and mean mass per unit area of any long-lived fall-back disk. Expressions for the torque components generated by interaction with the pulsar wind are obtained. It is not impossible that these could be responsible for small observable variations in pulse shape and spin-down rate but they are unlikely to be the source of the periodic changes seen in several pulsars. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, ?? -- ?? ()
Printed 19 December 2018
(MN LATEX style file v2.2)
Interaction and ablation of fall-back disks in isolated
neutron stars
P. B. Jones⋆
University of Oxford, Department of Physics, Denys Wilkinson Building, Keble Road, Oxford OX1 3RH, England
ABSTRACT
An analysis of ablation processes is made for a fall-back disk with inner and outer
radii external to the neutron-star light cylinder. The calculated ablation rate leads,
with certain other assumptions, to a simple expression relating the inner radius and
mean mass per unit area of any long-lived fall-back disk. Expressions for the torque
components generated by interaction with the pulsar wind are obtained. It is not
impossible that these could be responsible for small observable variations in pulse
shape and spin-down rate but they are unlikely to be the source of the periodic changes
seen in several pulsars.
Key words: accretion, accretion disks - stars: neutron - pulsars: general
1
INTRODUCTION
A simple model for neutron-star disk formation from super-
nova fall-back was described nearly twenty years ago (Michel
1988). More recently, there has been a revival of interest in
the subject in connection with topics such as planetary for-
mation (Lin, Woosley & Bodenheimer 1991; Wolszczan &
Frail 1992; Miller & Hamilton 2001), pulsar spin evolution
(Menou, Perna & Hernquist 2001a; Blackman & Perna 2004
), the anomalous X-ray pulsars (Chatterjee, Hernquist &
Narayan 2000; Marsden et al 2001; Ek¸si & Alpar 2005; Ek¸si,
Hernquist & Narayan 2005), and a common framework for
the anomalous X-ray pulsars, soft gamma-ray repeaters and
dim isolated thermal neutron stars (Alpar 2001).
Fall-back disks in isolated neutron stars may differ from
the accretion disks (α-disks) of binary systems in a number
of respects. It is possible for them to be at radii beyond the
light cylinder radius RLC , with temperatures observable in
the infra-red resulting from heating by X-rays and by the
pulsar wind rather than mass transfer and viscous evolu-
tion. A significant fraction of the mass may be in the form
of dust grains. Computations of the optical and infra-red
emission expected from fall-back disks have been made by
a number of authors (Foster & Fischer 1996; Perna, Hern-
quist & Narayan 2000), and there have also been several
experimental searches for the infra-red excess characteristic
of dust grains (see Bryden et al 2006). In this way, upper
limits of the order of 10−2M⊙ have been obtained for disk
masses in a number of radio pulsars (Phillips & Chandler
1994; Lohmer, Wolszczan & Wielebinski 2004).
The recent observation of an excess in the case of the
⋆ E-mail: [email protected]
anomalous X-ray pulsar 4U 0142+61 (Wang, Chakrabarty
& Kaplan 2006) has been interpreted by these authors as
evidence for an X-ray heated fall-back disk. The estimated
mass of the disk is quite small, Md ∼ 10−5M⊙, and a fac-
tor favouring its observation is the high X-ray luminosity
(∼ 1036 erg s−1) of the central neutron star relative to the
typical radio pulsar. Its estimated present inner radius is
well outside the light cylinder, ri ≈ 4.7RLC . An alternative
analysis by Ertan et al (2007) fits the complete infra-red and
optical spectrum of 4U 0142+61 by the emission of a viscous
gaseous disk with ri < RLC and an extinction to that object
identical with the estimate made independently by Durant
& van Kerkwijk (2006a). This type of disk is much more
complex than the one considered in the present paper and
should properly be regarded as forming a boundary con-
dition for the magnetosphere (see, for example, Cheng &
Ruderman, 1991). The ablation processes considered in this
paper would remain relevant at r > RLC for this type of
disk but their use to define the inner radius ri, as in Section
2.4, would not be possible.
Details of the fall-back disk formation are not described
here. Initially, the disk must have been internally ionized and
viscous, with mass and angular momentum transfer rates
intrinsic to the formation process. We assume that the es-
sential ideas of the evolutionary path described by Menou,
Perna & Hernquist (2001b) are valid. The thermal ioniza-
tion instability, in which free electron recombination causes
a sudden decrease in opacity, occurs first at the outer disk
radius but then propagates very rapidly inward. We accept
the conclusion of Menou et al that a neutral, passive, gaseous
disk is then formed, with no obvious source of internal vis-
cosity. The power input to the disk, considered here in Sec-
tions 2.1 and 2.4, is not adequate to maintain temperatures
2
P. B. Jones
above the thermal ionization instability temperature and so
does not materially affect the onset and propagation of in-
stability except, possibly, to delay it.
To summarize, the present paper assumes that the fall-
back disks with which it is concerned are passive, thin and
have internal temperatures well below the ionization insta-
bility temperature. Hence, following Menou et al, there is
negligible internal ionization and the viscosity of α-disks is
not present. We shall be concerned only with the interac-
tion of the pulsar wind with the disk surface regions so that
it will be possible to neglect the complex evolutionary dy-
namics of neutral dust and debris disks (see, for example,
Frank, King & Raine 2002). There is certainly interaction
between neutral dust grains but to a good approximation
we can treat the disk as a set of annuli each having a ra-
dius r and Kepler velocity vK . Warping of the disk then
occurs, in general, if torque components are r-dependent. It
is assumed here that, in the general case, the disk plane at
formation has a finite uniform tilt angle β, but that the disk
angular momentum is more nearly parallel rather than an-
tiparallel with the neutron-star spin. We consider only disks
that are external to the light cylinder, ri > RLC , and regard
their effect on the pulsar magnetosphere as a partial termi-
nation of the pulsar wind rather than a boundary condition
to be imposed on the solution for the magnetospheric fields.
This limitation on ri means that the equivalent force on
the disk derived from the Lense-Thirring effect (see, for ex-
ample, Wilkins 1972) is negligible in comparison with those
found in Section 3.1. The Robertson-Poynting effect can also
be neglected. The reason is that the equivalent force on the
disk depends on its Kepler velocity, vK ≪ c, but the torque
components derived in Section 3.1 from the azimuthal part
of the pulsar wind are independent of this factor.
The present paper is addressed principally to three
problems. These are: (i) an analysis of disk ablation pro-
cesses and estimates of the ablation rate and of the infra-red
luminosity; (ii) calculation of the alignment and precessional
torques acting between disk and neutron star; (iii) the ques-
tion of whether or not disk alignment, counter-alignment or
precession could give rise to time-variable phenomena that
are observable over intervals of no more than several years.
The last of these has been prompted by the observation
of periodic changes in timing residuals and in average-pulse
characteristics (Stairs, Lyne & Shemar 2000; Shabanova,
Lyne & Urama 2001; Haberl et al 2006) of several pulsars
or isolated neutron stars that appear to be most simply ex-
plained by Eulerian precession (for a recent review of this
interpretation, we refer to Link 2006). If the existence of
Eulerian precession were established unambiguously, there
would be far-reaching consequences for our present under-
standing of neutron-star internal structure. For this reason,
we attempt in Section 4 to see if residues of fall-back disks
could exist in these cases with inner radii small enough to
give variability on the observed time-scales.
2 DISK IONIZATION AND ABLATION
2.1 Momentum densities in the vacuum solution
In order to study the interaction between a disk and the pul-
sar wind we shall use inertial frame cartesian coordinates in
which the neutron-star spin angular velocity Ω is parallel
with the z-axis. Where convenient, this frame is also repre-
sented by spherical polar coordinates r, θ, φ. The magnetic
dipole moment is µ = BoR3, where R is the neutron star
radius and Bo the equatorial surface field. It is at an angle
ξ with Ω. There has been much computational work on the
aligned neutron-star magnetosphere, but in the general case
with ξ 6= 0, the work of Spitkovsky (2006), based on rela-
tivistic force-free electrodynamics, seems to be the only pub-
lished solution. Thus our present investigation is based on
the Deutsch vacuum solution for the electromagnetic fields
of a rotating neutron star in the form given in recent papers
by Michel & Li (1999) and Ek¸si & Alpar (2005). We note
that there appears to be some disagreement about a small
number of near-field terms in this solution (see also Ferrari
& Trussoni 1973; Good & Ng 1985) but these are not of pri-
mary concern for the torque calculations made here. Michel
& Li also observe that the static radial electric field term,
which remains significant at the light cylinder, is in prin-
ciple undetermined because the total electric charge of the
star and magnetosphere is an unknown quantity dependent
on the details of formation. From the electric displacement
D and magnetic flux density B given by this solution in the
inertial frame at r > RLC it is simple to write down the
spherical polar components of the momentum density,
p =
1
4πc
D × B.
(1)
For the torque calculation described in Section 3.1, we re-
quire them averaged over the pulsar rotation period. They
are:
hpri =
µ2Ω4 sin2 ξ
8πc5r2
(cid:0)1 + cos2 θ(cid:1) ,
hpθi = 0,
hpφi =
µ2Ω cos2 ξ sin θ
6πc2r5
+
µ2Ω
4πc2r5 (cid:18)1 +(cid:16) rΩ
c (cid:17)
2(cid:19) sin2 ξ sin θ.
(2)
(3)
(4)
The first term in equation (4) is derived from the product
of two static irrotational field components and hence gives a
solenoidal contribution to hpφi. Taken at face value, this rep-
resents a purely azimuthal component of momentum density
whose presence in the near field of an aligned rotating mag-
netic dipole appears not implausible. Integration to obtain
the rate of outward transfer of angular momentum across a
sphere of large radius r gives the vacuum torque,
hΓvi = cr2Z 1
−1
d(cos θ)Z 2π
0
dφhr × pi,
(5)
of which the z-component is the spin-down torque,
2µ2Ω3
sin2 ξ,
(6)
3c3
hΓvzi =
with hΓvxi = hΓvyi = 0, as expected in the inertial frame.
Naturally, identical results are obtained by integrating the
time-averaged torque, formed directly from the Maxwell ten-
sor, over the same surface. (In a frame of reference corotating
with the star, the time-averaged components of Γv are all
finite and Γv ∝ (Ω × µ) × µ; see Davis & Goldstein 1970.
Thus the direction of Ω changes in this frame. Therefore,
in general, it also changes relative to the dipole moment µ
which may move in this frame during pulsar evolution in an
uncertain manner.)
The relation between wind momentum density and
torque expressed by equation (5) is no more than classical
mechanics and remains valid, with hΓvi replaced by the true
torque hΓi, whatever the degrees of freedom contributing to
pφ. Thus hpφi is necessarily finite in a physical pulsar wind,
in which the electromagnetic fields are loaded by a density of
relativistic charged particles. However, it is unclear whether
or not hpθi would remain zero in a real system: a finite value
would not contribute to hΓi, whose x and y components
would necessarily be zero in the inertial frame. The effect
of these terms on the interaction between disk and pulsar
wind does not seem to have been considered previously and
is significant in the case of thin dust and debris disks similar
to those found in the major planets or in pre-main-sequence
stars (see, for example, Beckwith et al 1990) which present
a very small profile to pr.
2.2 Composition of the pulsar wind
A pulsar luminosity, observed outside the light cylinder, can
be broadly divided into a wind component Lw and black-
body radiation from the neutron-star surface in the form of
an X-ray component Lbol. The wind luminosity, which can
be estimated from Ω and the observed spin-down rate Ω, is
further divided into a Poynting flux of electromagnetic fields
and a relativistic particle component, Lw = Lem + Lp. In
previous work on disk ablation by the pulsar wind, Miller &
Hamilton (2001) assumed Lp to be the larger wind compo-
nent, and of baryonic composition.
However, this view of wind composition appears to be
at variance with theories of particle acceleration at r < RLC
that are widely accepted to be at least qualitatively valid.
In the corotating frame of reference, there may exist finite
electric-field gaps in the open magnetic flux-line regions of
the magnetosphere. The total current flow through these
is limited by the Goldreich-Julian charge density, σGJ =
−Ω · B/2πc, where B is the magnetic flux density. Thus
the total rate of baryon loss estimated from the polar-cap
area πR3Ω/c intersected by open magnetic flux lines and
from the polar-cap Goldreich-Julian charge density can be
at most of the order of
µΩ2
ec
,
(7)
Rb =
where µ is the neutron-star magnetic dipole moment. But
the size of the finite-field gaps is limited by intense electron-
positron pair formation with the consequence that the
Goldreich-Julian current density σGJ c must contain both
ion and positron components. In polar-cap models, particles
are typically accelerated to energies of the order of 103 GeV
per unit charge, though greater energies may be reached
in outer-gap models. It is also widely accepted that, as a
consequence of the gap-limitation process, large numbers of
electron-positron pairs are formed in open magnetic flux-
line regions external to the gaps. Thus the total rate of pair
formation can be given as Rep = κRb, but even its order of
magnitude, κ ∼ 103−5, is not well known. The energies of
these pairs are several orders of magnitude lower than the
Neutron-star fall-back disks
3
typical gap energy of 103 GeV. It is obvious that the particle
luminosity, estimated in the region of the light cylinder from
equation (7) and from the gap energy, must be Lp ≪ Lw in
most cases. For example, from the ATNF Pulsar Catalogue
parameters given for PSR 1828-11 (Manchester et al 2005),
we find Rb = 8×1031 s−1 giving an estimated Lp ≈ 1.3×1032
erg s−1, to be compared with Lw = 3.6 × 1034 erg s−1. This
is consistent with the usual assumption (see, for example,
Melatos & Melrose 1996) that the Poynting flux luminosity
Lem is the larger component of the wind at the light cylin-
der. We shall assume that this is maintained at disk radii.
2.3 Ablation processes
In this paper, we assume that the presence of a partially
ionized disk at r > RLC , well outside any possible Alfv´en ra-
dius, has little effect on fields and particle fluxes at r < RLC
and can be seen as a partial wind termination rather than
a modified boundary condition. The sequence of processes
in disk ablation is described in this Section. We show first
that the thermal X-ray flux Lbol ionizes the outer regions of
the disk surface and so makes possible the conversion of the
Poynting flux Lem to ion or proton kinetic energy. Ablation
is then the result of several possible secondary processes,
including neutron production in the interaction of the accel-
erated ions and protons with disk nuclei.
It is necessary to find the conditions under which a disk
consisting of neutral hydrogen, atoms of mass mA,Z, and
dust grains of radius a has significant ionization. We shall
assume that the disk has small thickness h and that it is
internally neutral, with a temperature well below the ion-
ization instability temperatures found by Menou, Perna &
Hernquist (2001b) for metal-rich compositions. But the sur-
face of a thin disk must have a finite temperature owing to
its interaction with the particle component of the wind mo-
mentum density pφ. As a result, a diffuse low-density surface
region is present which is exposed to the X-ray component
Lbol, principally photons of average energy E ≈ 2kBTs de-
termined by the neutron-star surface temperature Ts. The
gravitational restoring force on an atom, acting toward the
plane of a thin disk, in terms of the Kepler angular frequency
ΩK , is mA,ZΩ2
K w at height w above the plane of the disk.
Thus the depth of the diffuse surface must be of the order
of
,
mA,ZΩ2
K
= r kBT
A,Z(cid:17)1/2
(cid:16)w2
typically ∼ 107 cm for hydrogen at T = 103 K and for
ΩK = 10−2 rad s−1. The mean life of a neutral hydrogen
atom in this region against photoelectric ionization is
(8)
τpe =
λ
mH
4πr2E
Lbol
,
(9)
where λ is the mass attenuation length for low-energy pho-
tons in hydrogen and mH is the hydrogen atom mass. For a
typical radio pulsar, E = 102 eV equivalent to Lbol = 8×1031
erg s−1 for a neutron star radius R = 106 cm. The Particle
Data Group (Yao et al 2006) give λ = 1.0 × 10−4 g cm−2
so that even at a radius r = 1011 cm, the mean life is very
short, τpe = 15 s. For higher atomic numbers Z, lifetimes
against ionization are of the same order.
The extent of dust grain sublimation is uncertain. Dust
4
P. B. Jones
grains exposed to Lbol have an estimated equilibrium tem-
perature Tbol = TspR/2r ≈ 1300 K, provided details of
grain composition, emissivity and shape are neglected. Sub-
limation rates may be significant at this temperature which
is also approximately equal to the K-band Wien's displace-
ment law temperature. But its dependence on Ts and r make
it impossible to give any general statement. This estimate
assumes a grain radius a > 10−4 cm such that almost all
incident photons interact. The mass stopping power for low
energy electrons has been tabulated by Seltzer & Berger
(1982). It is so large that there is no doubt that the en-
ergy of the emitted photo-electrons is contained within dust
grains of these radii. Smaller grains, with incomplete reten-
tion of photo-electron energy, have lower radiative equilib-
rium temperatures. On the basis of the Dulong and Petit
law, the thermal equilibrium temperature is reached within
times short compared with the Kepler period of the disk.
The ionization time τpe is also smaller than the Kepler pe-
riod leading to the conclusion that, for interesting intervals
of Lbol and r, matter in the very diffuse surface regions of a
disk has a high degree of ionization independent of the state
of the dust component. Neutron star blackbody photons
therefore have an important effect on the state of the disk
but their energies, E ∼ 102 eV, are several orders of mag-
nitude too small to eject protons (a proton with 102 eV/c
momentum has negligible kinetic energy; see also Miller &
Hamilton 2001).
Following the discussion of the particle component of
Lw in Section 2.2, we shall assume that the fields at the
disk radius, though not necessarily at radii that are many
orders of magnitude greater, satisfy the ideal magnetohydro-
dynamic condition cE = −v0 × B, where v0 is the particle
velocity (see, for example, Melatos & Melrose 1996; for a
review of pulsar wind nebulae, see Gaensler & Slane 2006).
To the extent that this is true, the further conditions E < B
and E · B = 0 are also satisfied. In this case, the qualitative
details of particle acceleration as a result of the interaction
of the Poynting flux in the wind with the diffuse ionized disk
surface can be analyzed easily by making a Lorentz trans-
formation of velocity
v′
c
=
E × B
B2
(10)
from the inertial frame of Section 2.1 to a frame with fields
E′, B′ in which E′ = 0 (see, for example, Landau & Lifshitz
1962). In this frame, B′ is parallel with B and of magnitude
B′ = √B 2 − E 2. A proton formed by photoelectric dissoci-
ation of neutral hydrogen has an initial velocity (neglecting
Kepler and thermal velocities) in this frame of −v′ and its
subsequent orbit is a circle of radius v′/ωB in a plane perpen-
dicular to B′, with angular frequency ωB = ev′B′/cq′ where
q′ is its momentum in this frame. Thus its time-averaged
velocity in the original inertial frame is simply v′. Although
v′ and v0 are not, in general, exactly parallel, it is broadly
correct to say that the proton is transported forward, with
the existing particle component of Lw, into the body of the
disk. As this process continues, the Poynting flux is nec-
essarily converted to kinetic energy. The energy transfer is
linearly dependent on rest mass and thus is almost entirely
to protons or partially ionized atoms. The penetration depth
is very roughly defined by equating Lem/4πr2c with the par-
ticle pressure, as for the collisionless shock present in pulsar
wind nebulae (Rees & Gunn 1974).
In this elementary analysis, the time-dependence of the
E and B fields in the inertial frame has been ignored. This
appears a reasonable approximation because, at disk radii
r ∼ RLC , the proton ωB is several orders of magnitude
greater than Ω. We emphasize that the diffuseness of the
disk surface is the important factor in its interaction with
the Poynting flux component of the pulsar wind.
Disk ablation, which we define as baryon loss, can occur
principally through three processes. The initial assumption
made here is that the most important is the production of
neutrons in strong interactions between the accelerated pro-
tons or with heavier nuclei. The nuclear interaction cross-
section for baryons is defined primarily by the nuclear ra-
dius. Thus the mass attenuation length is λb ≈ 100 g cm−2.
A second source of loss is the possibility that, with the re-
flection of E and B fields that is neglected here, protons and
ions may be transported outward and away from the surface.
Finally, the secondary protons formed in relativistic proton
interactions inside the disk can also acquire a momentum
component parallel or antiparallel with B which allows them
to move out of the disk in a time small compared with the
neutron star rotation period. These processes are all inde-
pendent of the state of the dust component in the disk.The
order of magnitude ablation rates obtained from the first
of them are very different from those of Miller & Hamilton
(2001) which, apart from a solid angle factor, were simply
equal to Lw divided by the gravitional potential energy of a
disk proton.
Neutron production in strong interactions in the body of
the disk has characteristics similar to those found by Agosteo
et al (2005) who have measured the neutron spectra given
by the interaction of 40 GeV/c protons with a copper target.
(That the nuclear charge Z = 29 exceeds the disk average
is unimportant for this purpose.) There are two groups of
neutrons. Qualitatively, forward-directed neutrons interact
with disk nuclei, as do secondary protons and mesons, to
produce a hadronic cascade with length scale determined by
λb. The more numerous neutrons are from decay of target
nuclei and are emitted isotropically with average energy ∼ 3
MeV. These neutrons interacting with disk nuclei are either
moderated by elastic scattering or undergo (n, γ) capture.
The photon component of the cascades, mostly from π0 me-
son decay, can be a source of neutrons through excitation of
the giant dipole state in nuclei, but this is unimportant com-
pared with direct strong-interaction production. Provided
the disk is optically thick, the photon energy passes to ther-
mal degrees of freedom through Compton scattering and
photoelectric absorption. Very qualitatively, we can assume
that, in a high-energy cascade, protons of momentum less
than 1 GeV/c do not interact further but come to rest after
losing energy by ionization in accordance with the Bethe-
Bloch formula (see Yao et al 2006). For these considerations,
it does not matter whether the target nuclei in the body of
the disk, where the E and B fields are much reduced by
conversion of the Poynting flux to kinetic energy, are free or
still bound as atoms in dust grains.
For disks that are optically thick for high energy pro-
tons, satisfying ρh ≫ λb, where ρ is the matter density, dif-
fusion of the low-energy group of neutrons backward to the
surface is the important mechanism of ablation. Obviously,
this process is dependent on the angle χ between pφ and the
disk outward normal unit vector n⊥. The neutron produc-
tion rate grows approximately exponentially with depth, the
scale length being ∼ λb sin χ, but the probability of diffusion
to the surface without capture decreases more rapidly. It ap-
pears that only the primary and a small number of secondary
target nuclei in a cascade are effective sources of neutrons.
On this basis, and the neutron production measurements of
Agosteo et al, we shall adopt a number N , in the interval
1 < N < 10, as the mean number of neutrons escaping the
disk per proton accelerated by the incident Poynting flux
as described above. We shall also assume a constant value
Ep = 10 GeV for the accelerated energy.
Neutron diffusion is more probable from disks that are
not optically thick to high-energy protons, for example,
those with ρh ∼ 10λb. Thus the ablation rates of these can
be higher, as noted by Miller & Hamilton (2001). There is
also the interesting possibility in these instances that the net
momentum transfer normal to the disk may be antiparallel
with to the normal component of hpφi. The arguments for
this will be given in Section 3.3 with a brief discussion of
some consequences that might be observable.
It must be obvious that the analysis of ablation pro-
cesses given here is intended to be no more than qualitative
in the extreme. In particular, there is no rigorous solution for
the interaction of the E and B fields with the diffuse outer
regions of the disk. One difficulty is that the value of B at
the light cylinder varies by four orders of magnitude for the
specific neutron stars considered here. Thus the description
of Poynting flux conversion to particle kinetic energy given
here is more obviously valid for the smaller values of the
orbit radii v′/ωB than for the larger values present in long-
period neutron stars such as 4U 0142+61. But these failings
can be viewed less seriously when it is remembered that the
properties of the physical pulsar wind, even in the absence
of a disk, are not well known. The present analysis attempts
no more than to show that, on interaction with the disk,
much of the Poynting flux is irreversibly converted to pro-
ton kinetic energy and that the average proton momentum
is roughly parallel with the Poynting vector.
2.4 The ablation rate and disk luminosity
We consider first the effect of the time-averaged azimuthal
pulsar wind component hpφi on the disk. The coordinate
system adopted is that defined in Section 2.1. and the disk
is assumed initially plane, with uniform tilt angle β and rec-
tilinear line of nodes in the xy-plane at an angle γ with the
x-axis. For integration over the area of the disk it is conve-
nient to transform from spherical polar coordinates θ, φ to
an azimuthal angle ψ, defined in the disk plane and with
respect to the line of nodes. The angular relations satisfied
by points on the disk are:
cos χ = − cos(φ − γ) sin β,
cos θ = sin β sin ψ,
sin(φ − γ) =
cos(φ − γ) =
cos β sin ψ
cos ψ
p1 − sin2 β sin2 ψ
p1 − sin2 β sin2 ψ
,
.
(11)
Neutron-star fall-back disks
5
The power input per unit area of disk is −c2hpφ · n⊥i. For a
disk annulus of radius r and width δr, the total power input
to both sides is,
0
dψc2 hpφ · n⊥i .
δW = rδrZ 2π
It will be convenient to eliminate µ in favour of Lw = ΩhΓvzi
using equations (2), (4) and (6). Given that
Z 2π
dψ sin θ cos(φ − γ) = 4,
(12)
(13)
0
equation (12) can be re-expressed as,
3c3Lw sin β
δW = rδr
2πΩ3 sin2 ξ
3r5 cos2 ξ +(cid:18) 1
(cid:18) 2
r5 +
Ω2
r3c2(cid:19) sin2 ξ(cid:19) .
(14)
The first term on the right-hand side of this equation derives
from the solenoidal component of hpφi which was mentioned
in Section 2.1. But its singularity at ξ = 0 is clearly an
artefact of the unphysical vacuum solution for the E and B
fields and for disks external to the light cylinder with radii
ri < r < ro, we need retain only the final term in equation
(14). Integration gives the total estimated power input to
the disk from the azimuthal wind component,
3
2π
RLCLw sin β(cid:16) 1
ri −
W =
The rate of baryon loss per unit area of disk at radius r is,
ro(cid:17) .
(15)
1
3RLCNLw sin β
.
(16)
4π2r3Ep
B =
Evaluation of equations (15) and (16) for the estimated disk
parameters of 4U 0142+61 (Wang et al 2006) is of interest.
The wind luminosity derived from the ATNF parameters
(Manchester et al 2005) is 1.2 × 1032 erg s−1 and the inner
and outer radii are 4.7RLC and 16RLC , respectively. The
total power input to the disk is then W = 9 × 1030 sin β
erg s−1. At the inner radius, the power input per unit area
is 5 × 107 sin β erg cm−2 s−1 which is large in relation
to the gravitational potential energy per unit area of disk
divided by the t0 ∼ 105 yr lifetime of the neutron star,
GM ¯Σ/t0ri ∼ 5× 106 erg cm−2 s−1, where M is the neutron
star mass and ¯Σ ≈ 1.6 × 104 g cm−2 is the mean mass per
unit area found from the disk mass and radii given by Wang
et al. Thus the inner edge power input from the wind is large
compared with any possible rate of thermal energy release
by viscous evolution of the disk. Assuming re-radiation ex-
clusively in the infra-red, this power input corresponds with
a blackbody temperature of T 6 960 K, which is lower than
the temperature of 1200 K found by Wang et al. This is not
unreasonable owing to the very large value of Lbol ≫ Lw
for this neutron star. We have assumed the disk thickness
to be small and uniform. But it is possible that it has some
radial dependence, possibly of the form h ∝ r9/7 which is
characteristic of point-source illumination of an α-disk (see,
for example, Pringle 1996), and that some of the power in-
put to the disk is from Lbol as in the model fit of Wang et
al. The model adopted by these authors must include some
assumption about the absolute thickness of the disk, in ef-
fect, the solid angle it subtends at the neutron star. This
6
P. B. Jones
does not appear to be stated, but it is also of interest that
their fit involves a very large X-ray albedo, ηd = 0.97. Fi-
nally, we have to accept that non-disk contributions to the
the infra-red luminosity may also be significant. It is possi-
ble that the recent observation of short time-scale downward
fluctuations in the K-band intensity of this object (Durant
& van Kerkwijk 2006b), not readily accommodated within
the passive disk model, may be evidence for these.
Evaluation of equation (16) at the present value of Ω
gives an estimated baryon loss rate, at the inner radius, of
3 × 1010 cm−2 s−1 for N = 10 and Ep = 10 GeV, which is
negligible in comparison with the mean disk density ¯Σ. This
is an essential result because the ablation rate must have
been considerably higher at earlier times following neutron-
star formation. It is more interesting to integrate the abla-
tion rate over the life of the star, assuming simple spin-down
given by equation (6), and to compare the total baryon loss
per unit area at the inner disk radius ri with ¯Σ. We have,
Z t
0 Bdt =
3cI
4π2r3 (cid:18)N sin β
Ep (cid:19) (Ω0 − Ω) ,
(17)
where I is the neutron star moment of inertia and Ω0 its spin
angular frequency at formation. This assumes that, at earlier
times, the disk extended inward to smaller radii but has
since suffered ablation to its present value of ri. It provides
an estimate of a combination of the unknown parameters
Ep and N in terms of quantities accessible to experimental
measurement,
N Ω0 sin β
Ep
=
¯Σ
4π2r3
i
3cImH
.
(18)
The comparison for 4U 0142+61 gives a value for the right-
hand side of equation (18) of 3.4 × 107 erg−1 s−1 or 5 × 104
rad s−1 GeV−1. Insertion of the values we adopted in Sec-
tion 2.3 for N and Ep then leads to Ω0 sin β = 5× 104 which
is obviously about two orders of magnitude too large. But
given the considerable uncertainties both in our parameter
values and in those estimated by Wang et al, the discrep-
ancy is not disturbing. If the disk parameters of Wang et al
are accepted, the comparison suggests smaller Ep and larger
N than the values adopted in Section 2.3. Possibly, the two
processes of baryon loss listed but not considered there may
be more important than we assumed. We have also neglected
any increase or decrease in disk radii caused by angular mo-
mentum transfer from the neutron star via the torque com-
ponent Γ⊥ given by equation (24) in Section 3.1.
Apart from the uncertainties in N and Ep, we emphasize
that there are at least two principal sources of uncertainty
in our calculation of W. Our treatment of the wind-disk in-
teraction in Section 2.3 is elementary and does not give any
estimate of the reflection of wind energy. There is also the
uncertainty in our assumption that pθ makes no contribu-
tion because its time-average vanishes. This may not be too
unsatisfactory for torque estimates, as in Section 3.1, but
may nevertheless introduce an error in calculating the total
power input.
3 DISK-PULSAR TORQUES
3.1 Calculation of the torque
Calculation of the torque components acting on the disk
follows that for δW given by equations (11) and (12). It
will be convenient consider an annulus of width δr and to
resolve the torque acting on it in the plane of the disk into
components δΓp and δΓa that are, respectively, parallel with
and perpendicular to the line of nodes. The components are,
δΓp = −cr2δrZ 2π
0
dψ sin ψhpφi cos χ
sgn(cos χ)(cid:17) ,
2
3
(cid:16)cos χ +
and
δΓa = cr2δrZ 2π
0
dψ cos ψhpφicosχ
sgn(cos χ)(cid:17) ,
2
3
(cid:16)cos χ +
(19)
(20)
formed from the momentum flux perpendicular to the plane
(neglecting any wind-energy albedo) and with the assump-
tion that the disk is optically thick and re-radiates photons
as a Lambertian surface. Use of the time-averaged azimuthal
momentum density in these expressions is valid because the
neutron star rotation period is several orders of magnitude
smaller than the Kepler period of the disk. Thus it is per-
missible to treat the annulus, though not the whole disk, as
a rigid body. Evaluation of the first of these integrals gives
a zero disk precession torque, δΓp = 0. This cancellation of
contributions from the four quadrants of ψ is a consequence
of the vacuum-field form of hpφi, as given by equation (4),
and it is by no means obvious that it would hold for a phys-
ical pulsar wind. Integration of equation (20) gives a finite
torque which acts to align the angular momentum δLd of
the disk annulus with the neutron-star spin,
δΓa = cr2δr(cid:18) 3cLw
8πΩ3 sin2 ξ(cid:19)(cid:16)4Ia sin2 β +
3r5 cos2 ξ +(cid:18) 1
(cid:18) 2
π sin β(cid:17)
c2r3(cid:19) sin2 ξ(cid:19) ,
r5 +
Ω2
2
3
in which the integral,
(21)
(22)
Ia = Z π/2
0
dψ
cos3 ψ
,
p1 − sin2 β sin2 ψ
is a slowly varying function of β, 0.66 < Ia < 0.79.
The torque component perpendicular to the plane of
the disk is,
δΓ⊥ = cr2δrZ 2π
0
dψhpφicos χ
(cos β sin θ + sin β cos θ sin(φ − γ)) .
Evaluation of the angular integral using equations (11) gives,
(23)
δΓ⊥ = cr2δr(cid:18) 3cLw
8πΩ3 sin2 ξ(cid:19)(cid:0)2β cos β(cid:0)1 + sin2 β(cid:1)(cid:1)
c2r3(cid:19) sin2 ξ(cid:19) .
(cid:18) 2
3r5 cos2 ξ +(cid:18) 1
r5 +
Ω2
(24)
We shall assume that the angular momentum of a fall-back
disk is likely to be more nearly parallel, than anti-parallel,
with the neutron star spin, in which case the interaction
increases δLd and moves the disk outward to larger radii.
3.2 The effect of torques on the neutron star
The motivation for considering this aspect of the interac-
tion is that the angular momentum L of the neutron star
can be several orders of magnitude smaller than that of the
disk. In the case of 4U 0142+61, for example, the disk pa-
rameters estimated by Wang et al, including its very small
mass Md ∼ 10−5M⊙, give Ld = 2.0 × 1047 g cm2 s−1 which
is between one and two orders of magnitude larger than L.
Consequently, the changes in direction of both L and Ld
can be significant. Thus the Euler equation for the neutron-
L = Γ, is
star angular momentum in the absence of a disk,
replaced by,
L + Ld = (cid:0)Γ − Γd(cid:1) + (Γd) ,
(25)
for the whole system, in which Γd = Γa + Γp + Γ⊥ is the
integrated torque acting on the disk, and Γ is the modi-
fied non-disk torque. For hypothetical disks with ri ≫ RLC ,
the terms in equations (21) and (24) that are significant
contributors to the torque would be those derived from the
long-range component of pφ in equation (4). But the vacuum
torque hΓvi can also be obtained from this term, as could
the true hΓi if the true pφ were known (see equation 5). This
suggests, but does not prove, that the first, bracketed, term
on the right-hand side of equation (25) approaches hΓi in
this limit, the disk torque being cancelled by terms in hΓi.
However, there is no reason to suppose that this cancella-
tion is exact at disk radii r ∼ RLC of physical interest where
the near-field terms in equation (4), (21) and (24) are not
negligible.
The terms in Γd, including any non-zero Γp, are all
finite in the inertial frame when averaged over the neutron
star period. Thus the direction of L must change during the
evolution of the system, but at a rate that is unlikely to
be of observational interest. To confirm that this is so, we
can consider PSR 1828-11 as an example and suppose that
L is at a small angle u with the total angular momentum
L + Ld and precesses about it with angular frequency Ωp.
The torque required is Γd ∼ LuΩp = 2 × 1037 g cm2 s−2 for
a 500 day precession with amplitude 10−2, several orders
of magnitude greater than the spin-down torque which is
Lw/Ω = 2.3 × 1033 g cm2 s−2 for this pulsar.
3.3 Alignment and precesssion of the disk
The initial evolution of a very thin disk with small ¯Σ and
Ld ≪ L can be obtained directly from equation (25), to
the extent that we can neglect the effect of changes in its
geometrical form on calculations of Γd. The solution is,
Ld = Γ⊥
Ld β = −Γa
Ld sin β γ = Γp.
(26)
From a hypothetical initial state of a plane disk with β 6= 0,
the alignment torque Γa reduces β at a radius-dependent
Neutron-star fall-back disks
7
initial rate given by
β = −
δΓa
δLd
= −
1
2πΣ(GM )1/2r5/2 (cid:16) 3Lw
8πΩ(cid:17)
(cid:16)4Ia sin2 β +
π sin β(cid:17) ,
2
3
(27)
in which only the long-range term in equation (21) has
been retained. The disk mass per unit area Σ in equation
(27) is the local value, not the mean ¯Σ. There is initially
no precession for the vacuum-field pulsar wind assumed in
Section 3.1. But the r-dependent alignment rate destroys
the planar form of the disk and therefore introduces a fur-
ther torque contribution derived from the radial momentum
flux chpri + Lbol/4πr2c. This is the torque concerned in the
Pringle instability (Pringle 1996; see also Petterson 1977,
Frank et al 2002). Initially, it is a contribution to the pre-
cessional component Γp, but further evolution of the shape
and motion of the disk is more complex. It is worth adding
that any finite torque component derived from equation (19)
as the consequence of a momentum density hpθi 6= 0 in a
physical wind, rather than the vacuum field solution used in
Section 2.1, would also induce precession.
In relation to disk alignment and precession, the inter-
esting question is whether or not these processes are suffi-
ciently rapid to produce time-varying phenomena that are
observable within periods of no more than several years. For
phenomena involving the body of a 10−5M⊙ disk, the an-
swer must be in the negative because Γd is perhaps an order
of magnitude smaller than the spin-down torque Γvz yet, as
we have noted in Section 3.2, the disk angular momentum
Ld is at least of the same order as L.
But the inner edge of a disk must have small local val-
ues of Σ owing to its continual ablation by baryon loss, given
by equation (16). In principle, the alignment rate given by
equation (27) for these radii could become observably fast.
There is also the likelihood, mentioned briefly in the penul-
timate paragraph of Section 2.3, that for some intervals of
small Σ, the net momentum transfer normal to the disk is
antiparallel with the normal component of hpφi.
We can see how this arises by referring back to both the
description of hadronic cascade development in Section 2.3
and the torque calculations of Section 3.1. The integrands
in equations (19) and (20) both contain the rate of momen-
tum transfer normal to the disk and per unit area. Both
expressions are based on two assumptions; that the disk re-
radiates as a Lambertian surface and that it does so from
the surface of incidence with which the wind interacts. The
second assumption is certainly valid if the disk is optically
thick for the photon spectrum emitted and also has suffi-
cient thickness, at least of the order of Σ ∼ 102λb cos χ,
that the hadronic cascades terminate much nearer the sur-
face of incidence rather than the far surface. But the transfer
of energy in a cascade from the primary particle to thermal
degrees of freedom and to low-energy photons of perhaps
102 − 104 eV occurs predominantly in the late stages of its
development. Thus for Σ ∼ 10λb cos χ, the wind heats the
far surface of the disk, not the surface of incidence. The ef-
fect of this is to change the sign of the sgn(cos χ) terms in
equations (19) and (20) and, unless χ is relatively small, to
change the sign of δΓa, transforming it from an alignment to
a counter-alignment torque. There is a second factor having
the same effect. The production of low-energy neutrons re-
8
P. B. Jones
ferred to in Section 2.3 also occurs mainly in the late stages
of cascade development. For these Σ, the baryon loss rate
per unit area much exceeds that given by equation (16) and
is mainly through the far surface of the disk. This is sig-
nificant because massive low-energy particles are effective
carriers of momentum.
Validation of these qualitative arguments would require
very extensive Monte Carlo calculations of cascade develop-
ment. But although these have not been made, we suggest
that there is a strong likelihood that a counter-alignment
torque acts on the inner edge of a disk. Again, we shall not
attempt to consider the complex and detailed evolution of
a disk under these torque components, but restrict our dis-
cussion to the question of whether or not they could cause,
in principle, time-varying phenomena that are observable.
With a sign reversal of δΓa, equation (27) gives a time
constant ta for exponential growth of β,
1
ta ≈
Lw
8πΩΣ(GM )1/2r5/2
i
.
(28)
Evaluation for 4U 0142+61, with the parameters of Wang et
al and a local density Σ = 300 g cm−2, gives ta ≈ 1.1×1013 s,
which is too long to be of observable interest. But the value
of Lw is small for this neutron star and the inner radius of
the disk is large. Under the assumption that, for example,
PSR 1828-11 has a disk external to its light cylinder we
find, given the ATNF parameters for this pulsar, that growth
times could be quite short, ta ≈ 7 × 106(ri/RLC )5/2 s.
4 CONCLUSIONS
We have found that the azimuthal component of the pulsar
wind momentum-density exerts an r-dependent torque on
a thin dust or debris disk external to the light cylinder.
Both the Poynting and relativistic particle components of
the wind are responsible for ablation of the disk. For disks
that are optically thick to relativistic protons, the power
input, which we assume determines the infra-red luminosity,
is given by equation (15). This result neglects any albedo,
principally, in this case, reflection of E and B fields at the
disk surface which could be estimated only by a much more
complete treatment of this problem than we have attempted.
The rate of ablation is given by equation (16) in terms of
the wind luminosity Lw and the cascade parameters Ep and
N for which we have only qualitative estimates. Under the
very loose assumption that Ω0 sin β is a universal constant
for all isolated neutron stars that are formed with dust or
¯Σ should
debris fall-back disks, equation (18) shows that r3
i
also be a universal constant. From the parameters given for
4U 0142+61 by Wang et al, its value is 1.3 × 1038 g cm.
It is obvious that our assumption of a common forma-
tion spin angular frequency Ω0 for all neutron stars should
not be taken too seriously. But if the 4U 0142+61 analysis
of Wang et al and our analysis of disk ablation are both
accepted, equation (18) makes it possible to infer what disk
parameters are possible in other isolated neutron stars and
pulsars. The conclusion we reach on this basis is that long-
lived disks of the kind considered here could be present in
many pulsars without exceeding published limits on infra-
red luminosity.
It has not been possible to answer with any certainty the
question of whether or not the torque components calculated
here could be responsible for any observable time-varying
phenomena. PSR 1828-11 is an interesting case of a pulsar
that appears to have periodic variations in timing residuals
and in pulse shape consistent with small-amplitude Eulerian
precession of its spin axis relative to coordinates fixed in
the star (Stairs, Lyne & Shemar 2000). If the inferred 500d
period were equated with the growth time ta estimated at
the end of Section 3.3 for counter-alignment of the inner
edge of a disk, the required radius would be ri = 2.1RLC but
the mean disk density would have to ¯Σ ≈ 2 × 109 g cm−2.
An almost identical value would be required in the radio
pulsar, PSR B1642-03, which also appears to have periodic
timing residuals (Shabanova, Lyne & Urama 2001). This
might be thought a high density for a nominally thin disk
but, assuming, for example, an outer radius ro = 2ri, the
mass is only Md = 1.5×10−4M⊙, well below observed upper
limits. In principle, movement of the ionized inner edge may
have the potential to produce small changes in the observed
radio-frequency emission pulse shape and in the pulsar spin-
down rate. But it is not obvious why the changes should be
periodic, as observed.
The more recent case of the isolated neutron star RX
J0720.4-3125 in which the inferred blackbody temperature
and certain other pulse characteristics have been interpreted
as varying sinusoidally is rather different (Haberl et al 2006;
but see also van Kerkwijk et al 2007 for an alternative
analysis). This neutron star with a period of 8.39 s and a
high magnetic field has some properties in common with 4U
0142+61. The small luminosity, Lw = 4.7 × 1030 erg s−1
(Manchester et al 2005), is such that, even for ri = RLC ,
the variation-time estimate obtained from equation (28) is
several orders of magnitude too long to be of interest.
ACKNOWLEDGMENTS
It is a pleasure to thank the referee, Professor M. A. Alpar,
for some very helpful comments on this paper.
REFERENCES
Agosteo S., et al., 2005, Nucl. Instr. Meth. Phys. Res. Sec.
B, 229, 24
Alpar M. A., 2001, ApJ, 554, 1245
Beckwith S. V. W., Sargent A. I., Chini R. S., Gusten R.,
1990, Astron. J., 99, 924
Blackman E.G., Perna R., 2004, ApJ, 601, L71
Bryden G., Beichman C. A., Rieke G. H., Stansberry G. A.,
Stapelfeldt K. R., Trilling D. E., Turner N. J., Wolszczan
A., 2006, ApJ, 646, 1038
Chatterjee P., Hernquist L., Narayan R., 2000, ApJ, 534,
373
Cheng K. S., Ruderman M., 1991, ApJ, 373, 187
Davis L., Goldstein M., 1970, ApJ, 159, L81
Durant M., van Kerkwijk M. H., 2006a, ApJ, 650, 1082
Durant M., van Kerkwijk M. H., 2006b, ApJ, 652, 576
Ek¸si K. Y., Alpar M. A., 2005, ApJ, 620, 390
Ek¸si K. Y., Hernquist L., Narayan R., 2005, ApJ, 623, L41
Ertan U., Erkut M. H., Ek¸si K. Y., Alpar M. A., 2007, ApJ,
657, 441
Neutron-star fall-back disks
9
Ferrari A., Trussoni E., 1973, Ap. & Sp. Sci., 24, 3
Foster R. S., Fischer J., 1996, ApJ, 460, 902
Frank J., King A., Raine D., 2002, Accretion Power in As-
trophysics, 3rd ed., Cambridge University Press
Gaensler B. M., Slane P. O., 2006, Annu. Rev. Astron.
Astrophys., 44, 17
Good M. L., Ng K. K., 1985, ApJ, 299, 706
Haberl F., Turolla R., De Vries C. P., Zane S., Vink J.,
M´endez M., Verbunt F., 2006, A&A, 451, L17
Landau L. D., Lifshitz E. M., 1962, The Classical Theory
of Fields, 2nd ed., Pergamon, Oxford, pp. 69-72
Lin D. N. C., Woosley S. E., Bodenheimer P. H., 1991,
Nature, 353, 827
Link B., 2006, A&A, 458, 881
Lohmer O., Wolszczan A., Wielebinski R., 2004, A&A, 425,
763
Manchester R. N., Hobbs G. B., Teoh A., Hobbs M., 2005,
Astron. J., 129, 1993
Marsden D., Lingenfelter R. E., Rothschild R. E., Higdon
J. C., 2001, ApJ, 550, 397
Melatos A., Melrose D. B., 1996, MNRAS, 279, 1168
Menou K., Perna R., Hernquist L., 2001a, ApJ, 554, L63
Menou K., Perna R., Hernquist L., 2001b, ApJ, 559, 1032
Michel F. C., 1988, Nature, 333, 644
Michel F. C., Li H., 1999, Phys. Rep., 318, 227
Miller M. C., Hamilton D. P., 2001, ApJ, 550, 863
Perna R., Hernquist L., Narayan R., 2000, ApJ, 541, 344
Petterson J. A., 1977, ApJ, 216, 827
Phillips J. A., Chandler C. J., 1994, ApJ, 420, L83
Pringle J. E., 1996, MNRAS, 281, 357
Rees M. J., Gunn J. E., 1974, MNRAS, 167, 1
Seltzer S. M., Berger M. J., 1982, Int. J. Appl. Radiat.
Isot., 33, 1189
Shabanova T. V., Lyne A. G., Urama J. O., 2001, ApJ,
552, 321
Spitkovsky A., 2006, ApJ, 648, L51
Stairs I. H., Lyne A. G., Shemar S. L., 2000, Nature, 406,
484
van Kerkwijk M. H., Kaplan D. L., Pavlov G. G., Mori K.,
2007, ApJ, 659, L149
Wang Z., Chakrabarty D., Kaplan D. L., 2006, Nature, 440,
772
Wilkins D. C., 1972, Phys. Rev. D, 5, 814
Wolszczan A., Frail D. A., 1992, Nature, 355, 145
Yao W. -M. et al, 2006, J. Phys. G., 33, 1
This paper has been typeset from a TEX/ LATEX file prepared
by the author.
|
astro-ph/9810376 | 1 | 9810 | 1998-10-22T23:46:44 | Integrated Stellar Populations of Bulges: First Results | [
"astro-ph"
] | We present first results from an on-going survey of the stellar populations of the bulges and inner disks of spirals at various points along the Hubble sequence. In particular, we are investigating the hypotheses that bulges of early-type spirals are akin to (and may in fact originally have been) intermediate-luminosity ellipticals while bulges of late-type spirals are formed from dynamical instabilities in their disks. Absorption-line spectroscopy of the central regions of Sa--Sd spirals is combined with stellar population models to determine integrated mean ages and metallicities. These ages and metallicities are used to investigate stellar population differences both between the bulges and inner disks of these spirals and between bulges and ellipticals in an attempt to place observational constraints on the formation mechanisms of spiral bulges. | astro-ph | astro-ph |
Integrated Stellar Populations of Bulges:
First Results
By S. C. T R A G E R1, J. J. D A L C A N T O N2,
AND B. J. W E I N E R1
1Carnegie Observatories, 813 Santa Barbara Street, Pasadena CA 91101
2Department of Astronomy, University of Washington, Box 351580, Seattle WA 98195-1580
We present first results from an on-going survey of the stellar populations of the bulges and inner
disks of spirals at various points along the Hubble sequence. In particular, we are investigating
the hypotheses that bulges of early-type spirals are akin to (and may in fact originally have been)
intermediate-luminosity ellipticals while bulges of late-type spirals are formed from dynamical
instabilities in their disks. Absorption-line spectroscopy of the central regions of Sa -- Sd spirals
is combined with stellar population models to determine integrated mean ages and metallicities.
These ages and metallicities are used to investigate stellar population differences both between
the bulges and inner disks of these spirals and between bulges and ellipticals in an attempt to
place observational constraints on the formation mechanisms of spiral bulges.
1. Introduction
Current thinking considers two major pathways to the formation of spiral bulges.
Simplistically, either the bulge formed before the disk ("bulge-first," e.g. van den Bosch
1998), or formed from the disk ("disk-first," e.g., Combes & Sanders 1981). Previous
studies have shown that bulges of big-bulge spirals (like M31) share at least some stellar
population properties with mid-sized elliptical galaxies. They fall along the Dn -- σ relation
(Dressler 1987) and the Fundamental Plane (Bender, Burstein & Faber 1992). Moreover,
Jablonka et al. (1996) and Idiart et al. (1996) find that bulges of spirals (as late as Sc)
fall along the Mg -- σ relation defined by early-type galaxies, suggesting that bulges share
a mass-metallicity relation with elliptical galaxies. Together with kinematical evidence,
these observations have led to the idea that mid-sized ellipticals accrete disks from some
leftover gas reservoir, forming spiral galaxies (see, e.g., Kauffmann, White & Guiderdoni
1993). Therefore, the stars in spiral galaxy bulges should be similar to the stars in
ellipticals (i.e. metal-rich, high [Mg/Fe]; e.g., Worthey, Faber & Gonz´alez 1992; Trager
et al. 1998b) and unlike the stars in spiral galaxy disks.
Kinematical and surface brightness observations of small-bulge spirals indicate that
their bulges share many properties with their disks: Many bulges in these spirals are
better represented by a shallower exponential profile than by the steep de Vaucouleurs
profile (e.g., de Jong 1996) and many of these bulges exhibit disk-like kinematics (see
Kormendy 1993 for an excellent review). Moreover, Peletier & Balcells (1996) and de
Jong (1996) find that the color difference between bulge and disk in an individual galaxy
is much smaller than the variation in colors across galaxies of a single Hubble type.†
These observations have led these authors and others to propose that bulges (of small-
bulge spirals at least) are formed from the stars already present in their underlying disks.
That is, the stars in spiral galaxy bulges should be similar to the stars in spiral galaxy
† However, very recent observations by Peletier & Davies (this meeting), using HST WFPC2
and NICMOS imaging of the Peletier & Balcells sample of early-type spirals, have shown that
bulges are uniformly red, suggesting that the bulges are uniformly old and metal-rich. Clearly
these issues are by no means settled with current observational data.
1
2
S. C. Trager et al.: Integrated Stellar Populations of Bulges
disks (e.g., ages and metallicities of their inner disks, solar [Mg/Fe]) and unlike stars in
ellipticals.
These two scenarios do have testable consequences, and sophisticated techniques now
available can provide definitive answers to these questions. We have embarked on a multi-
year survey to probe the stellar content of the bulges and inner disks of spirals using
absorption-line strengths. These measurements can provide the critical tests needed to
understand the basic mechanisms driving bulge formation.
2. The stellar populations of spiral bulges
2.1. Sample and observations
We have selected a sample of 91 southern face-on spirals, ranging in type from S0/a to
Sdm, barred and unbarred, from the ESO (B) survey. These spirals are not interacting,
have major axes larger than 3′, are within 4000 km s−1 and are located at b > 20◦. The
observations have been made using the long-slit Boller & Chivens spectrograph at Las
Campanas Observatory in the blue, giving ≈ 2 A (σinst ≈ 35 km s−1) resolution in the
4000 -- 5200 A region. This spectral range covers portions of both the Lick/IDS (Burstein
et al. 1984; Worthey et al. 1994; Trager et al. 1998a) and Rose (1985, 1994) absorption-
line strength systems. To date, ten (primarily unbarred) spirals have been observed,
most along both major and minor axes, and data from four have been processed. Line-
strength profiles for these four galaxies have been calibrated onto the Lick/IDS system
using stellar observations in common with Jones (1996), and velocity dispersion profiles
and rotation curves have been derived following the Fourier-quotient procedure described
by Gonz´alez (1993).
2.2. The Mg -- σ relation
As described above, bulge-first formation mechanisms find support in the observation
that bulges in galaxies as late as Sc fall along the same Mg2 -- σ0 relation as early-type
galaxies (Jablonka et al. 1996; Idiart et al. 1996), suggesting the existence of a mass-
metallicity relation in spirals and a narrow spread in ages at fixed σ0. We confirm these
observational results: our bulges fall on the Mgb -- σ0 relation defined by early-type galaxies
(our observations do not cover the red sideband of the broad Mg2 index; see data in Trager
et al. 1998a). However, the Mg -- σ relation is inherently degenerate to compensating
variations in metallicity and age in old stellar populations -- large age spreads can exist
if there is a complementary age-metallicity relation (Worthey, Trager & Faber 1996;
Trager 1997). This intrinsic age-metallicity degeneracy in Mgb and Mg2 (and other metal
lines and broadband colors) prevents the Mg -- σ relation from being an effective stellar
population age discriminator. More sophisticated techniques are obviously necessary to
distinguish between bulge formation mechanisms.
2.3. Balmer-metal line diagrams
There is a way to break the age-metallicity degeneracy -- Balmer lines are more age-
sensitive than metal-line indices (with the possible exception of the G band; Worthey
1994). In the absence of of nebular emission or large numbers of blue horizontal branch
stars or blue stragglers, Balmer-line strengths reflect light-weighted mean turnoff temper-
ature of the composite stellar population -- i.e., the mean age of the population. Balmer
lines do have some intrinsic metal dependence, however, so a clean separation of age and
metallicity effects requires diagnostic diagrams combining Balmer line and metal line
strengths. Figure 1 presents such diagrams for the first four galaxies analyzed in our
sample.
S. C. Trager et al.: Integrated Stellar Populations of Bulges
3
Figure 1. Balmer -- metal-line diagrams for two Sa -- Sb spiral bulges (NGC 2775 and NGC 3054,
left panels) and two Sc bulges (IC 438 and NGC 1637, right panels). The Balmer line index HγA
is a sensitive measure of the presence of intermediate-age (1 -- 10 Gyr) stars in an old population.
The metal-line index C24668 is a sensitive measure of the bulk metallicity of an old population.
Although both indices are slightly sensitive to both age and metallicity, used together they can
effectively measure the mean age and metallicity in an old stellar population (Worthey 1994;
Worthey & Ottaviani 1997; Trager et al., 1998). Points are coded by the axis along which the
slit was placed, and grow smaller going out from the center. Closed symbols are as observed;
the open symbols are an attempt to correct for the emission fill-in in the HγA index using
an optimized stellar template (cf. Gonz´alez 1993). Lines represent models of Worthey (1994)
and Worthey & Ottaviani (1997): solid lines are contours of constant age and dotted lines are
contours of constant metallicity. The Sa/Sb bulges on the left are clearly quite old (10 -- 15 Gyr)
and metal-rich (metallicities as high or higher than solar) and are older than their disks by at
least a few Gyr. In contrast, the Sc bulges seem younger than the Sa -- Sb bulges (< 10 Gyr),
and the bulge of NGC 1637 is nearly as young as its disk. Both Sc bulges are also slightly
metal-poor (metallicities less than solar), suggesting that a mass-metallicity relation may exist
for spiral bulges, as suggested by the Mg -- σ relation.
The Balmer -- metal-line diagrams for large bulges (Sab -- Sbc, σ0 > 100 km s−1) suggest
that the bulges of these early-type spirals are consistent with having old, metal-rich
stellar populations, as seen in our own galaxy (t ∼> 10 Gyr, [C/H] ∼> 0; e.g., McWilliam
& Rich 1994; Bruzual et al. 1997). These bulges are also older than their inner disks by
several Gyr, suggesting that these bulges formed early on in the galaxies' histories, and
that bulge-first formation is a likely scenario for these galaxies.
On the other hand, small bulges (Sc, σ0 < 100 km s−1) seem to have younger and more
metal-poor populations (t ∼< 10 Gyr, [C/H] ∼< 0), and at least in some galaxies (NGC
1637), the bulge and inner disk are roughly the same age. Such an observation provides
strong support for a common origin of bulge and disk material -- that is, for a disk-
first formation scenario. However, contamination from emission is difficult to remove,
separation of bulge and disk light has not yet been attempted, and small amounts of
young or intermediate-age populations can significantly increase Balmer-line strengths
(Trager et al. 1998b). In these late-type, star-forming, small-bulged galaxies, such effects
may obscure truly old bulge populations.
4
S. C. Trager et al.: Integrated Stellar Populations of Bulges
3. Conclusions
The initial indications are that large bulges are genuinely old, metal-rich systems, as
expected from our own galactic bulge. The stellar populations of small bulges appear to
be younger and more metal-poor than the large bulges, but contamination from emission
lines and disk light make these conclusions uncertain for the moment. If these results
stand with more data and more extended analysis, we may find that bulges may have
more than one formation mechanism -- or even more than one mechanism may occur in
a single galaxy (Rix, this meeting).
We thank our collaborators Dr. R. O. Marzke and Dr. A. McWilliam for many stimulat-
ing converations during the course of this work. SCT would like to thank the organizers
for holding this very interesting workshop and for their financial support.
REFERENCES
Bender, R., Burstein, D., & Faber, S. M. 1992, ApJ, 399, 462
Bruzual, G., Barbuy, B., Ortolani, S., Bica, E., Cuisinier, F., Lejeune, T., & Schi-
avon, R. P. 1997, AJ, 114, 1531
Burstein, D., Faber, S. M., Gaskell, C. M., & Krumm, N. 1984, ApJ, 287, 586
Combes, F. & Sanders, R. H. 1981, A&A, 96, 164
de Jong, R. S. 1996, A&A, 313, 377
Dressler, A. 1987, ApJ, 317, 1
Gonz´alez, J. J. 1993, Ph.D. Thesis, University of California, Santa Cruz
Idiart, T. P., De Freitas Pacheco, J. A., & Costa, R. D. D. 1996, AJ,112,2541
Jablonka, P., Martin, P., & Arimoto, N. 1996, AJ, 112, 1415
Jones, L. A. 1996, Ph.D. Thesis, Univ. North Carolina
Kauffmann, G., White, S. D. M., & Guiderdoni, B. 1993, MNRAS, 264, 201
Kormendy, J. 1993, in Galactic Bulges, IAU Symposium 153, ed. H. Dejonghe & H. J. Habing,
p. 209
McWilliam, A. & Rich, R. M. 1994, ApJS, 91, 749
Peletier, R. & Balcells, M. 1996, AJ, 111, 2238
Rose, J. A. 1985, AJ, 90, 1927
Rose, J. A. 1994, AJ, 107, 206
Trager, S. C. 1997, Ph.D. Thesis, UC Santa Cruz
Trager, S. C., Worthey, G., Faber, S. M., Burstein, D., & Gonzalez, J. J. 1998a,
ApJS, 116, 1
Trager, S. C., Faber, S. M., Gonz´alez, J. J., & Worthey, G. 1998b, in prep.
van den Bosch, F. 1998, preprint (astro-ph/9805113)
Worthey, G. 1994, ApJS, 95, 107
Worthey, G., Faber, S. M., & Gonz´alez, J. J. 1992, ApJ, 398, 69
Worthey, G., Faber, S. M., Gonz´alez, J. J., & Burstein, D. 1994, ApJS, 94, 687
Worthey, G., Trager, S. C., & Faber, S. M. 1996, in Fresh Views on Elliptical Galaxies,
ed. A. Buzzoni, A. Renzini, & A. Serrano, A. S. P. Conf. Ser., vol. 86 (San Francisco: ASP),
p. 203
Worthey, G. & Ottaviani, D. L. 1997, ApJS, 111, 377
|
astro-ph/0007258 | 1 | 0007 | 2000-07-18T10:35:12 | Dynamical masses, time-scales, and evolution of star clusters | [
"astro-ph"
] | This review discusses (i) dynamical methods for determining the masses of Galactic and extragalactic star clusters, (ii) dynamical processes and their time-scales for the evolution of clusters, including evaporation, mass segregation, core collapse, tidal shocks, dynamical friction and merging. These processes lead to significant evolution of globular cluster systems after their formation. | astro-ph | astro-ph |
Massive Stel lar Clusters
ASP Conference Series, Vol. 211, 12, 2000, in press
A. Lan¸on, C.M. Boily, eds.
Dynamical masses, time-scales, and evolution of star
clusters
Ortwin Gerhard
Astronomisches Institut, Universitat Basel,
Venusstr. 7, CH-4102 Binningen, Switzerland
Abstract. This review discusses (i) dynamical methods for determining
the masses of Galactic and extragalactic star clusters, (ii) dynamical pro-
cesses and their time-scales for the evolution of clusters, including evap-
oration, mass segregation, core collapse, tidal shocks, dynamical friction
and merging. These processes lead to significant evolution of globular
cluster systems after their formation.
1.
Introduction
The Milky Way and probably all large galaxies contain old globular cluster
populations (see the review by Harris 1991). These old star clusters have an
approximately log-normal luminosity function, and the mean cluster luminosity
is somewhat brighter than MV = −7 with little dependence on the host galaxy
luminosity. In the Milky Way their typical mass-to-light ratios are M /LV ≃ 2,
and typical total masses are ∼ 2 × 105 M⊙ (Pryor & Meylan 1993). It is widely
assumed that the globular clusters we see today must be the part of an initially
larger population that survived the internal and external dynamical processes
leading to cluster destruction (e.g., Ostriker 1988).
One of the exciting results from HST has been the discovery of young star
clusters in starburst and interacting galaxies. Whitmore & Schweizer (1995)
found many hundreds of young clusters in the Antennae galaxies. Young cluster
systems have now been discovered in other interacting and merging galaxies,
in barred and starburst galaxies, and even dwarf starburst galaxies (e.g., ESO
338-IG04, Oestlin, Bergvall & Roennback 1998). The luminosity functions of the
young clusters are not log-normal, but seem to be better described by power-
laws, about ∝ L−2 . Carlson et al. (1999) use population synthesis models to
determine the ages of the blue clusters in the young merger remnant NGC 3597.
Based on these models they argue that the difference in the observed luminosity
function when compared to the Galactic globular clusters cannot simply be
an age effect, even if the young clusters formed with an intrinsic age spread.
Are these young cluster systems then a good model for what the Milky Way’s
globular cluster population could have looked like at birth?
This review gives a brief discussion of dynamical methods to determine
masses of distant and nearby star clusters (Section 2). It then goes on to de-
scribe a number of dynamical processes and their time-scales which will lead
to evolution and potentially destruction of star clusters over long time-scales.
1
2
Finally, the results of some evolution calculations for globular cluster systems
are briefly summarized (Section 3).
2. Dynamical Mass Determination for Star Clusters
In this Section we discuss methods for estimating star cluster masses from struc-
tural and kinematic measurements. Mass estimates based on stellar population
properties are discussed in U. Fritze von Alvensleben’s article in these proceed-
ings.
2.1. Virial Masses
(1)
A simple global mass estimate for a star cluster can be obtained from the virial
theorem. This says that, in equilibrium, the radius of a stellar system is propor-
tional to GM /V 2 , where M is the total mass and V the rms three-dimensional
velocity of the stars. The constant of proportionality generally depends on the
stellar density profile, but Spitzer (1969) showed that if the relation is expressed
in terms of the half–mass radius rh , this dependence is weak and the constant is
approximately 0.4 for realistic cluster profiles. If we furthermore assume that the
cluster is spherical, V 2 = 3σ2
k , where σk is the one-dimensional rms velocity dis-
persion along the line-of-sight, and write σ10 = σk/10 km s−1 and r5 = rh/5 pc,
then
k rh/G = 8.7 × 105σ2
MV = 7.5σ2
10 r5 M⊙ .
When using this formula to estimate star cluster masses from observed velocity
dispersions and radii, a few points should be noted:
(i) Because the virial mass (1) is a global estimate, it is independent of ve-
locity anisotropy. For example, shifting some stars to radial orbits while keeping
the (spherical) potential fixed, will result in a larger central velocity dispersion
but also lead to reduced velocities in the cluster halo. To maintain virial equi-
librium these changes must add in just such a way that the global σk remains
the same.
(ii) The dynamical evolution of star clusters leads to mass segregation and
the formation of a halo of low-mass stars on preferentially radial orbits (see
§3.2. below). For evolved clusters the measured half-light radius will therefore
in general underestimate the half-mass radius, the observed velocity dispersion
will underestimate the rms velocity dispersion, and eq. (1) will underestimate
the mass.
(iii) Sometimes only the velocity dispersion for stars in the core is known,
or the velocity dispersion from integrated light within some aperture. In these
cases, a dynamical model is needed to convert this to the rms σk . This intro-
duces some uncertainty in the mass estimate because the derived σk depends on
anisotropy.
With high-resolution spectra and HST photometry virial masses can be
determined for some young ‘superclusters’ seen in starburst galaxies. Masses for
three clusters in M82 are compared by Smith & Gallagher in these proceedings,
spanning a range from 3 × 105 M⊙ − 2 × 106 M⊙ .
2.2. Core Masses
3
9σ2
0
2πGI0 rc
Rood et al. (1972) gave a formula that is often used to determine core masses.
This is based on the dynamics of King models (King 1966; Binney & Tremaine
1987) and assumes that the velocity distribution in the core is isotropic:
(cid:18) M
L (cid:19) =
Here σ0 is the central velocity dispersion, I0 the central surface brightness and
rc the core radius. The product of the two last quantities is rather insensitive
to errors caused by seeing. Eq. (2) is very accurate as long as the assumption of
isotropy is met (Richstone & Tremaine 1986); but it can overestimate the mass
by a factor ∼ 2 if the system is actually radially anisotropic (Merritt 1988).
Core mass-to-light ratios can only be determined for well-resolved Galactic
star clusters for which the core parameters rc , I0 and σ0 can be estimated. Even
with the resolving power of HST the cores of distant young clusters cannot be
resolved.
.
(2)
2.3. Masses from Model Fitting
An alternative method of estimating cluster masses is fitting the photometric
and kinematic data with dynamical models. A simple such scheme was used
by Djorgovski et al. (1997) in their study of the M31 globular clusters mass-to-
light ratios. They used structural and photometric parameters for these clusters
obtained with HST and kinematic measurements in a rectangular aperture ob-
tained with Keck and HIRES. They then estimated an aperture correction from
King models to transform their measurements to central velocity dispersions,
and used a formula analogous to eq. (1) to estimate masses with the constant
again determined from models.
For some Galactic globular clusters large velocity samples are available and
in such cases much more detailed model fitting is possible (Pryor et al. 1989).
The masses of the Galactic globular clusters referred to in §1. (Pryor & Meylan
1993) have been determined by these techniques. A recent such study is Cot´e
et al. (1995) who investigated the dynamics of the globular cluster NGC 3201,
using a CCD surface brightness profile and a sample of 857 measured stellar
radial velocities to trace the velocity dispersion profile to large radii.
In such work the data are fitted by single- or multi-mass King-Michie mod-
els. In the multi-mass models, a power-law mass function for the cluster stars
is typically assumed, and for each mass bin mi , a distribution function of King-
Michie type (Michie 1963, Binney & Tremaine 1987) is used:
fi (E , J ) ∝ e−βJ 2 (cid:16)e−AiE − 1(cid:17) .
Here E , J are the specific energy and angular momentum of a star in the (spher-
ical) star cluster potential, and β can be thought of as specifying an anisotropy
radius. In the core of the cluster, stars of different masses are assumed to be in
equipartition (§3.3.), so that Ai ∝ mi . In the fitting procedure the free param-
eters are the radius, velocity and luminosity scale, the cluster’s concentration
parameter, the anisotropy radius, and the index of the mass function. These
(3)
4
parameters are determined from fitting to the measured surface brightness and
velocity dispersion profile. This leads to a determination of the mass-to-light
ratio profile and anisotropy profile, rather than just a single M /L as for the
previously described techniques. However, the fit is non-unique in the sense
that adding even fairly large numbers of faint low-mass stars in the halo (ex-
pected there from mass-segregation and evaporation, see §3.2. below) have little
effect on the observed profiles. In their study, Cot´e et al. (1995) find a steady
rise in M/L with distance from the cluster center, as expected from dynamical
evolution theory, and a global M /LB ≃ M /LV ≃ 2.0 ± 0.2.
2.4. Masses from Proper Motions
For nearby Galactic globular clusters, it is possible to measure stellar proper
motions in addition to radial velocities. Proper motion measurements give in-
formation about the velocity dispersions in two directions on the sky (radial
along pro jected R, and tangential), and for a spherically symmetric cluster they
are therefore in principle sufficient to determine the velocity ellipsoid as a func-
tion of radius, and thus the mass profile free of assumptions about anisotropy.
The pro jected proper motion dispersions σR (R) and σT (R) are related to the
intrinsic velocity dispersions σr (r) and σt (r) by Abel integral equations and
can thus be inverted (Leonard & Merritt 1989). Moreover, from the inferred
σr (r) and σt(r) one can predict the line-of-sight velocity dispersions σk (R) and
compare with independent radial velocity data. This provides a check on the
modelling and also can be used to determine the cluster distance. In terms of
global velocity dispersions, hσ2
k i = (hσ2
R i + hσ2
T i)/2 for a spherical cluster and
the correct distance.
In an early study along these lines Leonard et al. (1992) investigated radial
velocity and proper motion data for the globular cluster M13. They concluded
that the mean anisotropy of this cluster β = 3(σ2
R − σ2
T )/(3σ2
R − σ2
T ) ≃ 0.3 and
that the effect of the anisotropy on the mass determination is ∼ 20%. Much
more detailed modelling will be possible with the large proper motion surveys
currently in progress.
2.5. Non-Parametric Cluster Mass Distributions
With large samples of stellar velocities at hand, radial velocities or proper mo-
tions, it is possible to infer the mass distribution of the cluster without making
specific assumptions like King-Michie stellar distribution functions. This re-
quires solving the Jeans and pro jection equations for the intrinsic density and
velocity dispersions under some smoothness constraint, given the data. I do not
give the equations here, but refer to the papers mentioned below.
When the data consist of several hundred radial velocities, some assumption
about the anisotropy is still needed. Gebhardt & Fisher (1995) describe such a
non-parametric analysis of radial velocity data for four Galactic globular clusters,
assuming isotropy of the stellar orbits. With a few hundred stellar velocities in
each case the results are still noisy, but indicate radially increasing M /L-profiles
as expected. Merritt, Meylan & Mayor (1997) describe a similar analysis of the
cluster ω Centauri, assuming that it is oblate and seen edge-on, and that it is
described by a meridionally isotropic two-integral model. They find that the
5
mass distribution cannot be strongly constrained by their data, but appears to
be slightly more extended than the luminosity distribution.
As discussed above, proper motion data result in two independent velocity
dispersions in the plane of the sky, and thus, within a spherical model, they con-
tain sufficient information to determine the anisotropy of the stellar orbits. With
sufficiently large data sets it will therefore be possible to model the anisotropy
profile and mass distribution of a spherical cluster non-parametrically.
3. Dynamical Evolution Processes and their Time-Scales
3.1. Relaxation
On dynamical time-scales large star clusters (N >> 100) evolve collisionlessly.
That is, for some time after birth they are described by a quasi-equilibrium
phase-space distribution function (df) which is a function of the integrals of
motion (or of the stellar orbits) in the mean field potential. On longer times-
scales, however, the graininess of the distribution of stars becomes important,
and the dynamical evolution is no longer collisionless. Over a relaxation time
two-particle interactions then deflect the cluster stars from the orbits they would
otherwise have followed in the mean gravitational potential.
In the approximation of a homogeneous distribution of equal-mass stars,
with density ρ0 and isotropic Maxwellian velocity distribution with dispersion
σ0 , the two-body relaxation time is (Spitzer & Hart 1971)
(4)
yr
tr0 = 0.34
σ3
0
G2 m∗ ρ0 ln Λ
∗,⊙ (ln Λ)−1
4 m−2
10 n−1
= 1.8 × 108 σ3
10 yr.
Here n = 104 n4 pc−3 is the number density of stars, σ0 = 10 σ10 km s−1 the
velocity dispersion, m∗ = m∗,⊙ M⊙ is the mean mass per star, Λ is of order the
number of stars, and (ln Λ)10 = (ln Λ)/10. Note that tr,0 is inversely proportional
to the stellar phase space density. Central relaxation times in globular cluster
cores evaluated with eq. (4) are ∼ 107 − 109 yr.
The relaxation time often varies by large factors between the central and
outer parts of a stellar system. It is then useful to define a half-mass relaxation
time. For a virialized star cluster this is obtained from eq. (4) by replacing ρ0
with the mean density inside the system’s half-mass radius rh = 5r5 pc and σ2
0
by one third of the rms V 2 , and then using the virial theorem to express V 2
through rh and the total cluster mass M = 105M5 M⊙ . The result is (Spitzer &
Hart 1971)
ln 0.4N r3
GM !
0.14 N
h
where N is the total number of stars in the system and
5 M −1/2
3/2
h /GM )1/2 = 8.3 × 105 r
td ≡ rh/V = 1.58 (r3
5
is the dynamical time. For comparison with the local formula eq. (4), the fiducial
values used in eq. (5) correspond to a one-dimensional virial velocity dispersion
σ ≃ 3.4 km s−1 and a mean density n ≃ 96 pc−3 .
1/2
td = 7.2 × 108 M
5
3/2
∗,⊙ (ln Λ)−1
5 m−1
r
10 yr,
N
26 log 0.4N
(5)
(6)
trh =
1
2
=
6
On short time-scales cluster evolution is still collisionless, so long as td ≪ trh
(requiring N ≫ 100); for example, during the violent relaxation at formation.
The resulting quasi-equilibrium df subsequently evolves slowly in response to
collisions, which will tend to drive the system towards an isothermal energy
distribution. One aspect of such slow evolution would be a decrease of ellipticity
with dynamical age (Fall & Frenk 1985). This could be the reason for the
significantly rounder globular clusters in M31 and the Milky Way as compared
with the LMC and SMC clusters (Han & Ryden 1994).
3.2. Evaporation and Core Collapse
Collisions between single stars modify the stellar df in two ways. The rarer
process is ejection, in which a single close encounter leads one of the stars to
acquire a velocity greater than the local escape velocity ve and to escape from
the cluster. The time-scale for this is tej ≡ −N/(dN/dt) ≃ 1.1 × 103 ln 0.4N trh
∼ 104 trh (H´enon 1969). The more important process of evaporation is caused
by the cumulative effect of many weak encounters, which gradually increase a
star’s energy until v ≥ ve . It is easy to show that the rms escape velocity of
the cluster is just twice its rms virial velocity. Thus, on average, a particle with
v ≥ 2V = √12σk will escape. For a Maxwellian velocity distribution, a fraction
ǫ ∼ 0.74% of stars have v ≥ 2V ; these stars will escape in one dynamical time,
after which the high-velocity tail is repopulated only in ∼ trh . Thus one expects
the evaporation time scale of the cluster to be ∼ (ǫ/trh )−1 ; detailed calculations
(Spitzer & Thuan 1972) show that
(7)
tev ≡ − N (dN/dt)−1 ≃ 300 trh .
Because the evaporation is dominated by weak encounters, escaping stars leave
the cluster with only very small positive energy; thus the total energy of the
remaining cluster is nearly constant, but must be shared among a shrinking
number of stars. In virial equilibrium N 2/rh ≃ const. and thus ρ ∝ N/r3
h ∝ 1/N 5
3/2
h /M 1/2 ∝ N 7/2 . So as the cluster becomes denser, evaporation
and trh ∝ N r
accelerates and the system contracts to negligible mass and radius in finite time.
This evaporation model, however, neglects the fact that the evolution of the
stellar cluster is not homologous and that the rate of evolution is much faster in
the dense core than in the system’s outer parts. Stars gaining energy towards
evaporation build up an extended halo where the time scale for further energy
gain increases strongly, so that these stars may not in fact escape during the
age of the cluster. On the other hand, the dense core loses stars to the halo on
the much faster central relaxation time, and may collapse to very high densities
before Mtot and rh can change much.
This phenomenon of core collapse may be understood as a consequence of
the fact that self-gravitating star clusters have negative specific heat (Lynden-
Bell & Wood 1968):
In virial equilibrium the total energy E = −T , where
T is the total kinetic energy, which is proportional to the virial temperature
T /M = V 2 . As energy is withdrawn from the cluster, its kinetic energy increases
and so does the virial temperature. Since rh ≃ GM 2 /(2E ), the cluster thereby
contracts. Vice-versa, an energy production mechanism (e.g., from binary stars)
causes the cluster to cool and expand. Now the dense core of the cluster may be
7
,
,
(8)
approximately regarded as a virialized system in thermal contact with the rest
of the cluster. It is normally hotter than its surroundings and therefore loses
energy to them through stellar encounters. As a result it shrinks and becomes
yet hotter, loses still more energy to the surrounding stars, and contracts to
formally zero radius in finite time.
For a single-mass star cluster the late stages of core collapse are self-similar
(Lynden-Bell & Eggleton 1980, Cohn 1980). As the core radius
rc ≡ 3σc/p4πGρc
shrinks, the central density ρc and velocity dispersion σc increase and the core
mass Mc decreases according to
c , ∝ r0.77
c )1/2 ∝ r−0.11
ρc ∝ r−2.23
σc ∝ (ρc r2
Mc ∝ ρc r3
c
c
c
until the core radius and mass formally shrink to zero at time tcc . Moreover, the
density profile of the cluster outside the collapsing core has the same exponent:
ρ ∝ r−2.23 for rc(t) ≪ r ≪ rc (0).
The time-scale for core collapse is proportional to the central relaxation
time; for a single mass cluster it is tcc ≃ 330 trc once the collapse is in the
self-similar phase (Cohn 1980, Heggie & Stevenson 1988). The total time until
core collapse in Cohn’s (1980) model is ∼ 16 half-mass relaxation times or ∼ 60
initial trc . As Goodman (1993) has emphasized, the former number depends
on the mass distribution of the cluster, and tcc/trh will be less than 16 for
clusters more centrally concentrated than Plummer models. By noting that
c ∝ r−1.89
r−1
c drc/dt ∝ t−1
rc ∝ ρc/σ3
, one can solve for the asymptotic time-
c
dependence of the collapse:
σc ∝ (tcc − t)−0.06 , Mc ∝ (tcc − t)0.41 .
ρc ∝ (tcc − t)−1.18 ,
rc ∝ (tcc − t)0.53 ,
(10)
In summary, the collapse of a single mass cluster occurs in two stages (Cohn
1980). The longer part of the evolution is an evaporative phase, during which
stellar collisions simultaneously populate a halo and make the core shrink and
become denser. Only towards the end does the evolution accelerate and enter
the gravothermal instability phase of self-similar collapse.
(9)
3.3. Equipartition, Mass Segregation, and Multi-Mass Core Collapse
When the cluster contains different stellar masses, energy can flow not only from
the core to the halo, but also between stars of different masses. Stellar collisions
drive the system towards equipartition of energy mi hv2
i i = const. As eq. (4)
shows, relaxation proceeds faster for more massive stars. The equipartition
time-scale measures the rate at which a group of heavy stars with masses m2
loses energy to lighter stars of mass m1 (Spitzer 1969):
teq =
2 i)3/2
1 i + hv2
(hv2
8(6π)1/2G2m1m2n1 ln Λ
= 1.2
m1
m2
tr0 (m1 )
(11)
1 i = hv2
where we have assumed equal temperatures hv2
2 i and used eq. (4). Ini-
tially, hv2
i i is independent of stellar mass; thus the massive stars lose kinetic
8
energy and sink to the center, while lighter stars gain kinetic energy in collisions
and move outwards, a process called mass segregation. Moreover, eq. (11) shows
that mass segregation of the massive stars occurs before relaxation of the cluster
as a whole becomes significant.
However, equipartition may never be reached. A simple case considered by
Spitzer (1969) is one with two mass groups such that the heavy stars are much
more massive than the light stars, m2 ≫ m1 , but the total mass in the cluster
core is dominated by the light stars: M2 ≪ ρ1 r3
c1 . In this case equipartition
causes the formation of a small subsystem of heavy stars (M2 , m2 ) in the core of
the distribution of lighter stars. Applying the virial theorem to the subsystem
of heavy stars gives
0.4 GM2
rh2
+
4π G ρc1
3
.
(12)
(13)
k2 r2
h2 ,
hv2
2 i ≃
where the first term describes the self-energy of the subsystem M2 and the second
term its interaction with the system of light stars (k is a constant of order unity).
Spitzer (1969) noticed that the right-hand-side of eq. (12) has a minimum when
regarded as a function of rh2 . An equilibrium can therefore exist only if hv2
2 i is
greater than this minimum, that is, assuming equipartition, if
m2 (cid:19)3/2
M2
c1 ≤ 4.0 k−1 (cid:18) m1
ρ1 r3
In other words, if its mass is too large, the subsystem of heavy stars remains a
dynamically independent stellar system with mean square velocity greater than
the equipartition value. It continues to lose energy to the lighter stars, becoming
denser and hotter, and evolving away from equipartition all the time (mass
stratification instability). Fokker-Planck calculations (Inagaki & Wiyanto 1984,
Cohn 1985) show that in the end the subsystem of heavy stars core collapses
independently from the cluster of light particles, just like a single mass system.
The evolution to core collapse with a spectrum of stellar masses has been
considered by Inagaki & Saslaw (1985) and Chernoff & Weinberg (1990). The
detailed evolution occurs in several phases: First, collisions trying to establish
equipartition of energy lead to mass segregation and the formation of a heavy
mass core. Then this core undergoes the gravothermal instability, i.e., contracts
while remaining hotter than the mean temperature of the system and conducting
energy outwards. This collapse accelerates towards core collapse, and finally
goes over into a single-component collapse which reaches formally infinite central
density. The time scale for this multi-mass core collapse evolution is faster than
that for core collapse in any single component cluster, typically a factor of a few
faster than for a cluster composed of the heaviest mass alone.
Deep in collapse, the density slopes of all mass groups mk are characterized
by separate power laws in the region where the heaviest component dominates
the potential, such that approximately (Cohn 1985, Chernoff & Weinberg 1990)
d ln ρk /d ln r ≃ −1.89 (mk /mu ) − 0.35,
where mu is the mass of the heaviest species in the cluster. The overall mass
profile is then not self-similar.
(14)
9
A multi-mass core collapse, however, may be strongly influenced by the
stellar evolution of the more massive stars. This has two main effects: First,
the mass loss from massive stars through winds may lead to an overall mass loss
from the cluster, and thus cause an adiabatic expansion. Secondly, the finite
stellar life-time tM S limits the time tcc during which they can core collapse, such
that tcc(m∗ ) ∼< tM S (m∗ ). Both effects greatly increase the overall core collapse
times; compared to a system of point masses within the range (0.4 − 15)m⊙ ,
Weinberg & Chernoff (1989) find an increase by about a factor of 30 − 60 in
their globular cluster models, including the expansion effects.
A reasonable approximation for the stellar lifetime of all but the most mas-
sive stars is tM S ≃ 9 · 109 (m∗/m⊙ )−2.6 yr. Thus if the most massive stars leave
black hole remnants of 3 M⊙ , these together with tight binaries will dominate
the evolution after 5 · 108 yr, while if the most massive remnants are 1.4 M⊙
neutron stars, they and the binaries will dominate after 4 · 109 yr. The latter
time scale approaches the time expected for core collapse in typical Milky Way
globulars.
3.4. Reversing core collapse
A number of energy source mechanisms can stop core collapse (e.g., Goodman
1993): (i) Processes that generate kinetic energy in the core directly, such as
binary stars transferring energy to the field stars in collisions. (ii) Mass loss
processes that heat the core indirectly above its virial temperature, including:
normal stellar evolution, accelerated stellar evolution by the formation of massive
stars in mergers, and ejection of stars through binaries. In all cases the net result
is adiabatic expansion and cooling of the core.
Only hard binaries contribute to field star heating. Binaries are hard if
their binding energy Eb = −Gm1m2/a (with a their semi-ma jor axis) exceeds
the mean kinetic energy, Eb > 3m∗σ2 ; those with Eb < 3m∗σ2 are called soft
binaries. Heggie’s law (Heggie 1975, Hut 1983) states that, on average, hard
binaries get harder by collisions with field stars, and soft binaries get softer.
Essentially, the orbital velocity of a hard binary is on average greater than the
velocity of an incoming field star, and the tendency towards energy equipartition
therefore results in a net transfer of energy to the field star. The opposite is
true for soft binaries, which gain net energy and eventually dissolve. The binary
behaves like a mini-system with negative specific heat: as energy is withdrawn
from it, the orbit shrinks, the orbital velocity increases, and the binary hardens.
When the binary becomes sufficiently hard, the typical recoil from a collision
with a single star becomes large, and the binary will eventually be kicked out
of the cluster. Just like in the Sun, the binaries providing the nuclear energy
source will eventually be ’burned’.
Binaries can be formed by a close gravitational interaction of three stars
(‘three-body binaries’), by dissipational tidal capture, or at the time of star
formation (‘primordial binaries’). To be effective in reversing core collapse,
binaries must have orbital semi-ma jor axes
a < Gm∗ /3σ2 = 3 σ−2
10 m∗,⊙ AU.
(15)
The formation of a hard three-body binary requires a close encounter be-
tween two stars (δv ≃ v) with a third star in the immediate vicinity, such that
10
one of the three stars acquires additional energy, leaving the other two as a
bound pair. Thus the time-scale is (Goodman 1984, Binney & Tremaine 1987)
t3 ≃ (np2 v)−1 (np3 )−1 ≃
σ9
n2 (Gm∗ )5 ≃ N 2
c ln Nc tr0 ,
(16)
where p ≃ Gm∗/v2 , v = O(σ) because low relative velocities dominate, and we
have used eq. (4) to express Goodman’s result in terms of the central relaxation
time and the total number of stars in the core, Nc . This implies that about
1/Nc ln Nc three-body binaries form per central relaxation time. In other words,
three-body binaries become important if the final core collapse is driven by fewer
than 100 of the largest mass stars.
3.5. Tidal field and tidal shocks
A steady tidal field lowers the energy threshold beyond which stars are no longer
bound to the cluster.
It thus increases the mass loss rates from evaporation,
both because the fraction of stars in the velocity distribution that escape in a
dynamical time increases, and because the decreasing number of stars in the
cluster leads to shorter relaxation times. The Quintuplet and Arches young
clusters (Figer et al. 1999) in the inner Galactic bulge are two clusters for which
these tidal effects are very important (Kim et al. 1999).
In reality, the tidal field is not stationary in the frame of the cluster stars.
This complicates the escape process, but more importantly it leads to a new dy-
namical process in cluster evolution, referred to as gravitational shocking (Os-
triker, Spitzer & Chevalier 1972). The tidal field acting on the cluster may
suddenly increase in strength when the cluster passes through the disk of its
host galaxy, or when it comes close to the high-density inner bulge near peri-
galacticon of its orbit.
In both cases, the perburbations to the stellar orbits
caused by the tidal shock lead to an effective energy input in the cluster which
makes the cluster less bound and accelerates mass loss from internal processes.
A detailed recent discussion of this process is given by Kundi´c & Ostriker
(1995) and Gnedin, Lee & Ostriker (1999). For stars in the outer parts of
the cluster, the tidal perturbation can be approximated as impulsive because
of the short time-scale of passage through the galactic disk.
In the cluster’s
central parts, on the other hand, the stellar orbital time-scales are short and
adiabatic invariance reduces the effects of the perturbation. Traditionally these
effects of the tidal shock were described by a first-order term h(∆E )ts i, which
denotes the net energy gain averaged over stellar orbits at a given position in the
cluster. Kundi´c & Ostriker (1995) noticed that the second-order term h(∆E )2
ts i
is typically even more important and competes with two-body relaxation near
the half-mass radius in driving evolution of the cluster’s internal structure. This
can speed up core collapse by a factor of three (Gnedin, Lee & Ostriker 1999).
Cluster destruction is accelerated; recent modelling of the evolution of the Milky
Way’s globular cluster system shows that the typical time to destruction becomes
comparable to the typical age of the Galactic globulars (Gnedin & Ostriker
1997).
11
3.6. Dynamical friction and merging
As already noted by Tremaine, Ostriker & Spitzer (1975), massive star clusters
experience dynamical friction against field stars as they move along their orbits
through the host galaxy. Because of the frictional drag the cluster loses orbital
energy and spirals into the galaxy center, where the tidal field becomes ever
stronger and will eventually dissolve the cluster.
The time-scale for dynamical friction for a cluster on a circular orbit at
initial radius ri = 2ri,2 kpc in a singular isothermal sphere with circular velocity
vc = 250 v250 km s−1 is (Chandrasekhar 1943, Binney & Tremaine 1987)
1.17r2
i vc
ln Λ GM
(17)
1
3
=
(18)
tdf =
= 46 r5 σ−1
10 v250 pc.
5 (ln Λ)−1
i,2 v250 M −1
= 2.64 × 1011 r2
10 yr
where M5 is again the cluster mass in units of 105 M⊙ . The friction time-scale
thus scales with the square of the cluster’s initial radius in the potential, and is
inversely proportional to its mass. It is the inner, most massive clusters which
are affected first.
If we continue to model the inner parts of the host galaxy as an isothermal
sphere with MG (r) = v2
c r/G and use the virial theorem [eq. (1)] for the cluster,
we can determine the radius at which the incoming cluster will dissolve as
rhvc√7.5 σk
rdis ≡ rh (cid:18) MG (rdis )
M (cid:19)
Young clusters formed in the high-density regions of starburst galaxies would
thus contribute to the build-up of the nuclear bulge after being dragged inwards
by dynamical friction and tidally shredded by the tidal field.
In some circumstances it may be possible that several young clusters are
born close enough to eachother to tidally interact and even merge. To quantify
this we use an approximate merging criterion fitted by Aarseth & Fall (1980)
to the results of N-body merger simulations. For the escape velocity of the
clusters at pericenter p of their relative orbit we take an approximate expression
h )1/2
k /(1 + p2/1.2r2
e (p) = 27.6σ2
assuming two overlapping Plummer spheres, v2
(see also the discussion in Gerhard & Fall 1983). Then the criterion for merging
becomes
6σk !2 1 +
+ vp
h !
4rh (cid:19)2
p2
(cid:18) p
∼< 1
1.2r2
where vp is the relative velocity at pericenter. Here we have used the virial
equation (1), and σk is again the one-dimensional rms velocity dispersion of the
cluster. For head-on collisions, this formula predicts merging for vp ∼< 6σk =
60 σ10 km s−1 (slightly more than √2 times the rms escape velocity from each
cluster), or ∆v = √36 − 27.6 σk ≃ 3σk ≃ 30 σ10 km s−1 for their relative velocity
at large separations.
It also shows that merging requires the two clusters to
come within several half-mass radii of eachother for merging to occur, at cor-
respondingly smaller approach velocities. The most likely situation for this to
happen would be when two clusters are born from the same giant molecular
cloud complex.
(19)
1
2
12
3.7. Evolution of globular cluster systems
The evolution of globular cluster systems has recently been modelled in a num-
ber of studies, among others by Gnedin & Ostriker (1997), Murali & Weinberg
(1997), Baumgardt (1998) and Vesperini (1998). These models combine as-
sumptions about the initial cluster mass function and cluster locations with
evolutionary models for individual clusters. In the models the various processes
described above are considered, and treated in some studies by parametrized
mass loss rates or analytic approximations to the results of N-body simulations,
in others as diffusion terms in Fokker-Planck models. Some of the main results
of these studies are:
(i) Globular cluster systems in galaxies evolve significantly. In the Milky
Way the typical cluster destruction time is of order the age of the system, and
about half of the present globulars will be destroyed in the next Hubble time.
(ii) Clusters in the inner regions of their host galaxy are disrupted most
rapidly. Similarly, clusters on eccentric orbits are preferentially destroyed over
clusters on tangential orbits.
(iii) Low-mass and high-concentration clusters are disrupted by evaporation,
loosely bound clusters and those on central or eccentric orbits by tides, and
massive inner clusters by dynamical friction and tides.
(iv) Low-mass clusters are destroyed most efficiently and initial power-
law mass distributions tend to become transformed towards approximately log-
normal mass distributions.
References
Aarseth S.J., Fall S.M., 1980, ApJ, 236, 43
Baumgardt H., 1998, A&A 330, 480
Binney J.J., Tremaine S., 1987, Galactic Dynamics. Princeton Univ. Press,
Princeton
Carlson M.N., et al., 1999, AJ, 117, 1700
Chandrasekhar, S., 1943, ApJ, 97, 255
Chernoff D.F., Weinberg M.D., 1990, ApJ, 351, 121
Cohn H., 1980, ApJ, 242, 765
Cohn H., 1985, IAU Symp. 113, Goodman J., Hut P., eds, 1985, 161
Cot´e P., Welch D.L., Fischer P., Gebhardt K., 1995, ApJ, 454, 788
Djorgovski S.G., et al., 1997, ApJ, 474, L19
Fall S.M., Frenk C.S., 1985, IAU Symp. 113, Goodman J., Hut P., eds, 1985,
285
Figer D.F., et al., 1999, ApJ, 525, 750
Gebhardt K., Fischer P., 1995, AJ, 109, 209
Gerhard O.E., Fall S.M., 1983, MNRAS, 203, 1253
Gnedin O.Y., Ostriker J.P., 1997, ApJ, 474, 223
Gnedin O.Y., Lee H.M., Ostriker J.P., 1999, ApJ, 522, 935
Goodman J., 1984, ApJ, 280, 298
13
Goodman J., 1993, ASP Conf. Ser., 50, 87
Han C.H., Ryden B.S., 1994, ApJ, 433, 80
Harris W.E., 1991, ARAA, 29, 543
Heggie D.C., 1975, MNRAS, 173, 729
Heggie D.C., Stevenson D., 1988, MNRAS, 230, 223
H´enon M., 1969, A&A, 2, 151
Hut P., 1983, ApJ, 272, L29
Inagaki S., Wiyanto P., PASJ, 36, 391
Inagaki S., Saslaw W., 1985, ApJ, 292, 339
Kim S.S., Morris M., Lee H.M., 1999, ApJ, 525, 228
King I.R., 1966, AJ, 71, 64
Kundi´c T., Ostriker J.P., 1995, ApJ, 438, 702
Leonard P.J.T., Merritt D., 1989, ApJ, 339, 195
Leonard P.J.T., Richer H.B., Fahlmann G.G., 1992, AJ, 104, 2104
Lynden-Bell D., Eggleton P.P., 1980, MNRAS, 191, 483
Lynden-Bell D., Wood R., 1968, MNRAS, 138, 495
Merritt D., 1988, AJ, 95, 496
Merritt D., Meylan G., Mayor M., 1997, AJ, 114, 1074
Michie R.W., 1963, MNRAS, 125, 127
Murali C., Weinberg M.D., 1997, MNRAS, 291, 717
Oestlin G., Bergvall N., Roennback J., 1998, A&A 335, 850
Ostriker J.P., 1988, IAU Symp. 126, Grindlay J.E., Davis Philip A.G., eds, 271
Ostriker J.P., Spitzer L., Chevalier R.A., 1972, ApJ, 176, L51
Pryor C., McClure R.D., Fletcher J.M., Hesser J.E., 1989, AJ, 98, 596
Pryor C., Meylan G., 1993, ASP Conf. Ser., 50, 357
Richstone D.O., Tremaine S., 1986, AJ, 92, 72
Rood H.J., Page T.L., Kintner E.C., King I.R., 1972, ApJ, 175, 627
Spitzer L., 1969, ApJL, 158, L139
Spitzer L., Hart M.H., 1971, ApJ, 164, 399
Spitzer L., Thuan T.X., 1972, ApJ, 175, 31
Tremaine S., Ostriker J.P., Spitzer L., 1975, ApJ, 196, 407
Vesperini E., 1998, MNRAS, 299, 1019
Weinberg M.D., Chernov D.F., 1989, in Dynamics of Dense Stellar Systems,
Merritt D., ed, Cambridge Univ. Press, Cambridge, 221
Whitmore B.C., Schweizer F., 1995, AJ, 109, 960
|
astro-ph/0512108 | 1 | 0512 | 2005-12-05T14:21:16 | The Space Interferometry Mission Astrometric Grid Giant-Star Survey. I. Stellar Parameters and Radial Velocity Variability | [
"astro-ph"
] | We present results from a campaign of multiple epoch echelle spectroscopy of relatively faint (V = 9.5-13.5 mag) red giants observed as potential astrometric grid stars for the Space Interferometry Mission (SIM PlanetQuest). Data are analyzed for 775 stars selected from the Grid Giant Star Survey spanning a wide range of effective temperatures (Teff), gravities and metallicities. The spectra are used to determine these stellar parameters and to monitor radial velocity (RV) variability at the 100 m/s level. The degree of RV variation measured for 489 stars observed two or more times is explored as a function of the inferred stellar parameters. The percentage of radial velocity unstable stars is found to be very high -- about 2/3 of our sample. It is found that the fraction of RV-stable red giants (at the 100 m/s level) is higher among stars with Teff \sim 4500 K, corresponding to the calibration-independent range of infrared colors 0.59 < (J-K_s)_0 < 0.73. A higher percentage of RV-stable stars is found if the additional constraints of surface gravity and metallicity ranges 2.3< log g < 3.2 and -0.5 < [Fe/H] < -0.1, respectively, are applied. Selection of stars based on only photometric values of effective temperature (4300 K < Teff < 4700 K) is a simple and effective way to increase the fraction of RV-stable stars. The optimal selection of RV-stable stars, especially in the case when the Washington photometry is unavailable, can rely effectively on 2MASS colors constraint 0.59 < (J-K_s)_0 < 0.73. These results have important ramifications for the use of giant stars as astrometric references for the SIM PlanetQuest. | astro-ph | astro-ph |
The Space Interferometry Mission Astrometric Grid Giant-Star
Survey. I. Stellar Parameters and Radial Velocity Variability
Dmitry Bizyaev1,2, Verne V. Smith1,3, Jose Arenas4, Doug Geisler4, Steven R. Majewski5,
Richard J. Patterson5, Katia Cunha1,6, Cecilia Del Pardo7, Nicholas B. Suntzeff8, Wolfgang
Gieren4
[email protected], [email protected], [email protected], [email protected],
[email protected], [email protected], [email protected], [email protected],
[email protected]
ABSTRACT
We present results from a campaign of multiple epoch echelle spectroscopy of relatively faint
(V = 9.5 − 13.5 mag) red giants observed as potential astrometric grid stars for the Space
Interferometry Mission (SIM PlanetQuest). Data are analyzed for 775 stars selected from the
Grid Giant Star Survey spanning a wide range of effective temperatures (Tef f ), gravities and
metallicities. The spectra are used to determine these stellar parameters and to monitor radial
velocity (RV) variability at the 100 m s−1 level. The degree of RV variation measured for 489
stars observed two or more times is explored as a function of the inferred stellar parameters.
The percentage of radial velocity unstable stars is found to be very high -- about 2/3 of our
sample. It is found that the fraction of RV-stable red giants (at the 100 m s−1 level) is higher
among stars with Tef f ∼ 4500 K, corresponding to the calibration-independent range of infrared
colors 0.59 < (J − Ks)0 < 0.73. A higher percentage of RV-stable stars is found if the additional
constraints of surface gravity and metallicity ranges 2.3 < log g < 3.2 and -0.5 < [Fe/H] <
-0.1, respectively, are applied. Selection of stars based on only photometric values of effective
temperature (4300 K < Tef f < 4700 K) is a simple and effective way to increase the fraction
of RV-stable stars. The optimal selection of RV-stable stars, especially in the case when the
Washington photometry is unavailable, can rely effectively on 2MASS colors constraint 0.59 <
(J − Ks)0 < 0.73. These results have important ramifications for the use of giant stars as
astrometric references for the SIM PlanetQuest.
Subject headings: stars: abundances, fundamental parameters, oscillations, late-type -- techniques: ra-
dial velocities
1National Optical Astronomy Observatory, Tucson, AZ,
85719
2Sternberg Astronomical Institute, Moscow, 119899,
Russia
3McDonald Observatory, University of Texas, Austin,
TX 78712
4Universidad de Concepcion, Concepcion, Chile
5University of Virginia, Dept.
of Astronomy, Char-
lottesville, VA 22903-0818
6Observatorio Nacional, Rio de Janeiro, Brazil
7University of Texas at El Paso, Dept. of Physics, El
1.
Introduction
The Grid Giant Star Survey (GGSS; Patter-
son et al. 2001) is a partially-filled, all-sky sur-
vey to identify giant stars from 9 . V . 17.5
mag using the Washington M, T2 + DDO51 filter
photometric prescription outlined in Majewski et
al. (2000). A primary motivation for the survey
and a driver of its design is the selection of stars
suitable for the Astrometric Grid of the Space In-
Paso, TX 79968
Chile
8Cerro Tololo InterAmerican Observatory, La Serena,
1
terferometry Mission (SIM PlanetQuest). These
Grid stars, which serve as astrometric references
against which the motions of SIM targets are mea-
sured, must themselves be as astrometrically sta-
ble as possible. Thus, they must be free of stellar
or significant planetary companions as well as at-
mospheric activity (spotting/flaring) that will in-
duce photocenter wobbles at the several microarc-
second level on of order a decade-long timescale
(approximately the duration of SIM). G and K
giants were selected by the SIM project as the pri-
mary stellar constituent of the SIM Astrometric
Grid because these stars are the most luminous,
common stellar type found in all directions of the
sky. High intrinsic luminosity places giant stars
at greater distances for a given apparent magni-
tude, and this increased distance decreases the an-
gular scale of any fixed linear astrometric wobble,
thus making these particular stars less likely to be
problematical as references. The GGSS project
has been undertaken with the specific aim of iden-
tifying subsolar metallicity giant stars, which are
intrinsically more luminous than solar metallicity
giants, and also less likely to have planets (see Fis-
cher & Valenti 2005, and references therein).
Despite consideration of such effects for care-
ful selection, all SIM Astrometric Grid candidates
require pre-mission monitoring to assess their like-
lihood of astrometric stability during the mis-
sion. One method is to search for variability in
other properties, such as brightness and radial ve-
locity (RV). Indeed, we are presently involved,
along with other groups, in a large campaign of
echelle high-resolution RV monitoring of Astro-
metric Grid candidates. Fortunately, not all vari-
ability in RV (or luminosity) necessarily translates
to detrimental astrometric variability. However,
some intrinsic photometric and/or RV variability
is due to processes that, although benign in terms
of photocenter wobble (e.g., radial pulsations), can
mask otherwise detectable signatures of problem-
atical sources of variability, like astrometric wob-
bles due to planetary, brown dwarf or stellar com-
panions. Therefore, because campaigns to monitor
radial velocity variability to the precision needed
for vetting SIM targets (i.e. ∼ 100 m s−1) are
both expensive and time consuming, it is useful
to understand how intrinsic RV variability of gi-
ant stars depends on other stellar properties (like
temperature, metallicity, and surface gravity) to
optimize efficient selection of giant stars for the
Astrometric Grid.
Although it has long been known that stars at
the top of the RGB are both brightness and RV-
variable (see e.g. Pryor et al. 1988; Cote et al.
1996), there has been little systematic study of
this variability as a function of stellar atmospheric
parameters, especially at fainter absolute magni-
tudes. Monitoring of radial velocities for bright
red giants has indicated a high probability of their
RV variations at the level of 100 m s−1 (see Hatzes
& Cochran 1994, and references therein). Joris-
sen et al. (1997) found that red giants with spec-
tral types late G to early K are stable and giants
with later spectral types are all variable. At the
same time, study of Hipparcos red clump giants
(Adelman 2001) reveals photometric stability in
this evolutionary phase. Radial velocity jitter at
large amplitude is expected for metal-poor, lumi-
nous red giants (MV ≤ -1.4) (see Carney et al.
2003).
In their assessment of the suitability of giant
stars as SIM Astrometric Grid members, Frink et
al. (2001) studied a proxy sample of 86 nearby gi-
ant stars selected from the Hipparcos catalogue.
Their study included three to eight epochs of high
precision (5-8 m s−1) echelle RV measurements on
timescales from days to a year. Frink et al. found
that most (73 of 84) of the giant stars they in-
vestigated have a stable RV at the level of < 100
m s−1, and the peak of the distribution of mea-
sured dispersions in velocity occurs at 20 m s−1.
However, Frink et al. do report that RV varia-
tion is more probable for redder stars ((B − V ) >
1.2 in their sample). While the closest match to
the survey presented here, the Frink et al. survey
contains stars that are much brighter than the ex-
pected SIM Grid sample, and, moreover, they did
not probe variability as a function of metallicity,
which is a key aspect of the stars being found in
the GGSS.
Our goal here is to carry out an initial RV-
stability assessment of stars more like those ex-
pected to fill the Astrometric Grid by focusing on
stars taken directly from the GGSS itself. The
hope is that RV stability may correlate with some
intrinsic stellar property, such as effective temper-
ature, surface gravity, or metallicity. Although
the precision of our RV measurements is an or-
der of magnitude less than that of Frink et al.
2
(2001), it is appropriate for finding RV wobbles
at the level needed to identify astrometrically-
detrimental grid candidates for the fainter, > 1
kpc distant GGSS candidates (and, indeed, is the
precision at which monitoring campaigns of SIM
Grid stars are being conducted). Moreover, our
sample is almost an order of magnitude larger than
that in the Frink et al. survey. The observations
studied and discussed here were obtained under an
initial JPL-sponsored program in which the inter-
nal RV accuracy was set at about 100 m s−1.
2. Observations and Data Reduction
2.1. Sample Selection and Biases
The GGSS finds giant stars by photometry in
the Washington M, T2 + DDO51 filters accord-
ing to the methods described in Majewski et al.
(2000). The (M − T2, M − DDO51) two-color
diagram (2CD) effectively distinguishes between
late type dwarfs and giants based on the surface-
gravity-sensitive Mgb+MgH feature near 5150 A
(where the 130A -- wide DDO51 filter is centered).
Photometry was obtained in 1302 evenly placed
fields across the celestial sphere, each of area 0.5 --
1.0 deg2. The 2CD (i.e., the Mgb+MgH feature)
is secondarily sensitive to metallicity and the po-
sitions of stars in the 2CD can therefore be used
to derive a crude estimate of the metallicity of the
likely giant stars. Metal-poor giants are the most
widely separated stars in the 2CD from the locus
of dwarf stars, and therefore these are the easiest
to identify with our methods. Additional details
regarding the GGSS can be found in Patterson et
al. (2001).
The selection of stars for the present study is
biased according to the same specific set of cri-
teria used in the original agreement of the GGSS
collaboration with the SIM Project regarding the
selection of Astrometric Grid candidates from the
GGSS.1 Giant star candidates are identified from
1In the past year, the SIM Project has modified somewhat
the criteria for the selection of stars for the Astrometric
Grid, with a focus on stars with magnitudes more typically
V ∼ 9 − 11 mag and chosen from both the GGSS and the
Tycho-2 catalogues in a ratio of approximately 1:3. The Ty-
cho stars are selected without foreknowledge of metallicity,
and so are expected to be typically near solar metallicity
and at smaller distances than stars from the GGSS. In this
respect, all of our results are still relevant, however, since
the stars included in the present study span the range of
the 2CD and assigned photometric metallicities
therefrom. From the apparent magnitude, color
and metallicity, an estimate of the absolute mag-
nitude and a photometric parallax distance is as-
signed (Rhee et al. 2001). For each field, all giant
candidates with V < 13.5 mag are rank-ordered
according to distance, with the most distant star
assigned highest priority. The top four ranked
stars in each field, yielding more than 4000 can-
didates of 9.5 . V . 13.5 mag, were passed
onto both medium and high resolution spectro-
scopic observing campaigns (the latter forming the
database explored here). Typically, though not
exclusively, the most distant giant candidates in
each field were also among those with the lowest
photometric metallicities brighter than V = 13.5
mag. Thus, our spectroscopic sample is biased
toward more metal-poor and distant stars than
would be found from a random selection of gi-
ant stars in the same magnitude range. From the
above described sample of more than four thou-
sand stars, we explore here a random subsample
of 775 stars in both celestial hemispheres, which
constitutes those stars for which multiple high-
resolution spectra have been obtained during the
first two years of the GGSS follow-up program.
2.2. Spectroscopy
Spectroscopic observations of 434 stars in the
northern subsample of GGSS candidates were con-
ducted in the period from January 2001 to De-
cember 2002. We made use of the 2.1m telescope
at the McDonald Observatory and the Sandiford
Cassegrain echelle spectrograph, which provides
R = 55000 resolution. Th-Ar comparison spec-
tra were taken right before every program star ob-
servation. The quality of the data was monitored
by observing one or two RV standard stars from
Nidever et al. (2002) per night. The adopted setup
of the spectrograph enabled us to cover the 5000-
5900 A spectral range and achieve a signal to noise
(hereafter S/N ) level of order 20-40 in reasonable
exposure times (10-30 minutes).
The observed spectra were reduced from two-
dimensional to "echelle" format in a standard way
with the IRAF2 software package. Corrections for
properties expected for both GGSS and Tycho-2 stars.
2IRAF is distributed by NOAO which is operated by AURA,
Inc. under contract with the NSF.
3
bias, flat field, and scattered light were applied,
and cosmic ray hits cleaned out by the IRAF's
tasks from crutil package. The internal accuracy
of the wavelength calibration via the Th-Ar lamp
spectra was on average 0.001 A (and not worse
than 0.0018 A), which corresponds to an RV ac-
curacy of order 55 m s−1 (and not worse than 100
m s−1).
Resampling the spectra introduces systematic
errors in the wavelength calibration that degrades
the RV accuracy. Thus, to preserve the wave-
length calibration the spectrum orders were not
concatenated to produce a single spectrum, but
rather were maintained individually at their nat-
ural sampling.
Southern hemisphere GGSS stars were observed
with the 1.2-m Swiss telescope and CORALIE ve-
locimeter in 2001-2004 at the ESO La Silla Ob-
servatory (Chile), as described by Arenas et al.
(2002). CORALIE is an improved version of the
ELODIE spectrograph (Baranne et al. 1996). The
effective resolution of CORALIE is 50000, it covers
the wavelength range 3870 -- 6800 A and it pro-
vides a precision of typically 30 m s−1 in the radial
velocities. Typically a S/N of 10 was achieved for
program targets. The radial velocity and its ac-
curacy comes directly from CORALIE's reduction
package TACOS (Baranne et al. 1996).
CORALIE provides RVS for 375 program stars
directly from the TACOS software. We found that
spectra S/N ≤ 4 have systematically low RV accu-
racy and rejected them from further analysis. The
whole sample of good S/N CORALIE candidates
is thus reduced to 341 objects.
While we cannot obtain abundances directly
from CORALIE spectra, we incorporate the RVs
of these stars in our analysis and adopt photomet-
rically estimated stellar parameters for these stars.
Since the southern sample was observed with a
smaller aperture telescope than the northern one,
it is biased toward brighter stars: the distribution
of the southern V-magnitudes peaks at 11.6 mag
whereas the northern sample distribution peak is
located at 12.7 mag.
3. Data Analysis
3.1. Radial Velocities
RVs from the McDonald spectra were measured
using a cross-correlation methodology with the
help of IRAF's fxcor task. After the primary
reduction the continuum of each order looks al-
most flat except at its edges where the continuum
level looks systematically lower and the S/N is de-
graded. For the RV analysis we discard the pixels
at order edges, which is about ∼ 20% of the total.
Each spectral order was cross-correlated against
the RV template separately. The template cho-
sen was the Arcturus spectrum (Hinkle et al.
2000). The cross-correlated pieces of the tem-
plate spectrum were convolved with a Gaussian
corresponding to the FWHM of the Th-Ar lines
found for each order.
In addition, the corre-
sponding piece of the calibrating Th-Ar spectrum
was cross-correlated with a laboratory Th-Ar tem-
plate (Palmer & Engleman 1983). The Th-Ar
template for each order was also convolved with
the same Gaussian as for the stellar template in
that order. The shift in the radial velocity be-
tween the observed Th-Ar and the laboratory ref-
erence frame was subtracted from the stellar RV
obtained from the cross-correlation via the tem-
plate. Thus we not only have taken into account
the varying instrumental profile by this order-by-
order treatment, but we are able to make multiple
estimates of the RV for a single observation and
thereby assess the errors in our velocities for each
star. A few of the spectral orders indicated sig-
nificantly different RVs from the others because
of poor wavelength calibration due to a sparse
Th-Ar line sample. We discarded such "bad" or-
ders from our analysis. The remaining orders' ra-
dial velocities (RVi) give us an independent as-
sessment of the mean RV (RV0) and its accuracy:
RV0 = Σ RVi · wi / Σ wi, where wi is the in-
verse square of the individual RV error provided
by IRAF. The dispersion of this estimate δRV was
found as δRV = pΣ wi · (RVi − RV0)2 / Σ wi.
We show an example of the order-to-order radial
velocity difference assessed for the program star
G1113+00.20 in Figure 1.
4
3.2. Stellar Parameters Derived from an
Automated Spectroscopic Analysis
In addition to the measured RV's, physical stel-
lar parameters can be derived for the red giants
studied here. The effective temperature (Tef f ) of
giant stars can be determined from their broad-
band colors. According to an observational cali-
bration given by Alonso et al. (1999), near-infrared
colors (J − Ks) can be used to estimate Tef f for
red giants, even with no information on the stellar
metallicity and surface gravity. The full release of
the All-Sky Point Source Catalog of the Two Mi-
cron All-Sky Survey (2MASS) makes available the
near-infrared colors of all GGSS candidates. Fig-
ure 2 shows the Tef f distribution of the combined
northern plus southern samples as estimated from
(J − Ks) colors according to the Alonso et al. pre-
scription. As may be seen, our candidates span a
range of temperatures characteristic of red giants.
The typical accuracy of 2MASS colors is 0.03 mag
in our sample magnitude range. It corresponds to
a 100 K accuracy in Tef f .
The internal accuracy of the Tef f calibration in
Alonso et al. (1999) from (J − Ks) is of order 125
K. On the other hand, Bessell, Castelli, & Plez
(1998) provide a calibration of the Tef f (J − Ks)
relation for giant stars from NMARCS stellar at-
mosphere models. The Bessell, Castelli, & Plez
results are reproduced by the Alonso et al. data
fairly well for typical values of the surface grav-
ity (see Figure 12 in Alonso et al. 1999). The
relation from Bessell, Castelli, & Plez (1998) is
metallicity-dependent but the difference between
stars of [Fe/H]=0.0 and -2.0 (which corresponds
to the bulk of our objects, see below) affects Tef f
by less than 100 K and would not have a signif-
icant effect on the applied calibration procedure.
From the previous two paragraphs, we expect ∼
160 K accuracy in our derived Tef f .
For the northern sample (434 stars) observed at
McDonald Observatory, the reduced echelle spec-
tra can be used to determine the surface grav-
ity, log g, and metallicity, [Fe/H], by comparison
to a library of synthetic stellar spectra. For this
purpose we produced artificial spectra using the
MOOG-2002 spectral synthesis package (Sneden
1973). The corresponding stellar models were
computed by the ATLAS 9 code (Kurucz 1993).
We computed a model set covering the 3800 --
5600 K temperature range, 0.2 ≤ log g ≤ 5.6, and
−3.0 ≤ [Fe/H] ≤ +0.5 with steps in these param-
eters of 200 K, 0.2 dex and 0.2 dex respectively.
In all the models we assumed a typical value for
the microturbulent velocity of 2 km s−1.
We selected three pieces of spectra to estimate
the stellar parameters. The Arcturus spectrum
(used as an RV template in §3.1) was split into
40 A pieces, and the optimal synthetic spectrum's
piece and corresponding model parameters were
found by least-square fitting. The stellar param-
eters of Arcturus (obtained by Griffin & Lynas-
Gray (1999)) were best reproduced in the regions
of 5175 -- 5215, 5215 -- 5255, and 5700 -- 5740 A.
These wavelength intervals used to derive the stel-
lar parameters are not simply dominated by Fe
absorption lines, but also by (principally) lines of
Mg I, Ti I, Cr I, Ni I, and to a lesser degree V I.
The abundances derived in this way are therefore
not direct "Fe abundances", but an overall "metal-
licity" defined by a mixture of both Fe-peak and
α elements; we refer to this abundance as sim-
ply metallicity and denote it in the remainder of
this paper as [Fe/H]. Although significantly non-
solar abundance ratios, relative to Fe, could create
additional noise in the metallicity estimates, the
vast majority of stars in the sample are well above
[Fe/H] ∼ -2. Within this metallicity regime only
Mg/Fe, and to a lesser extent Ti/Fe and Cr/Fe,
are mildly non-solar and typically enhanced. This
may introduce a small trend of slightly larger de-
rived overall metallicities as the Fe abundance de-
creases.
To match the wavelength intervals defined
above, spectra of program stars were corrected to
zero radial velocity incorporating the RVs defined
in § 3.1. The continuum level in the fragments of
spectra was fit by a linear function using the con-
tinuum task from the IRAF package. The pieces
of synthetic spectra utilized for the comparison
were convolved with a Gaussian corresponding to
the FWHM of each echelle order to reduce their
spectral resolution to that in the observed data.
Then we find the best-fit model for each piece by
the least-square method. The values and errors of
log g and [Fe/H] were finally obtained by averag-
ing these same derived parameters across the set
of three modeled pieces.
With this method, we derived the stellar pa-
rameters (Tef f , log g, [Fe/H]) for the 434 north-
5
ern GGSS stars (see Table 1). The [Fe/H] for the
whole GGSS sample has also been estimated pho-
tometrically (Rhee et al. 2001)3 and one can com-
pare the values derived by each method (photo-
metric and spectroscopic). Figure 3 shows that
the two [Fe/H] estimates for most stars of inter-
mediate metallicity ([Fe/H] = -1.0 to 0.0) follows
a linear relation with dispersion of the order of 0.4
dex, but systematically offset by 0.16 dex (in the
sense that the spectroscopic metallicities tend to
be larger). This small offset could be due partially
to mildly elevated Mg/Fe, Ti/Fe, or Cr/Fe ratios
in the metal poor stars. There are some extreme
outliers in Figure 3. Some of these stars have large
error bars, either in their spectroscopic or photo-
metric metallicities. Many of the outlying points
in Figure 3 correspond to the hottest stars in our
sample. These stars occupy the border region be-
tween giants and dwarfs on the giant-dwarf dis-
crimination diagram from Majewski et al. (2000),
and their stellar parameters determined from pho-
tometric data may be biased.
Figure 4 shows the relation between the surface
gravity and effective temperature for the northern
hemisphere sample. Most of the stars in Figure
4 span ranges typical for red giants. Here we see
the success rate of the GGSS for photometrically
identifying bona fide red giants to be extremely
high: more than 98 % of the stars that were photo-
metrically identified by the Majewski et al. (2000)
method to be giant stars have spectroscopic grav-
ities supporting this characterization.
3.3. Errors in Atmospheric Parameters
In order to analyze the accuracy of the derived
stellar parameters as a function of the quality of
observed spectra, we performed Monte Carlo sim-
ulations deriving the basic stellar parameters for
a set of selected synthetic spectra deteriorated by
varying amounts of added noise. We considered
a set of signal-to-noise ratios from 3 to 30 which
covers the typical range of S/N in our observed
spectra. Besides random noise, we also added
variations of Tef f and RV into the model spec-
tra, distributing them uniformly within ±100 K
and ±100 m s−1 range from the true value. Fi-
3These photometrically derived metallicities are those that
have been delivered by the GGSS collaboration to the SIM
project.
nally, we assumed that the microturbulence veloc-
ity might take an arbitrary value in the range 1 to
2 km s−1. The stellar parameters were defined for
a hundred resulting synthetic spectra over the typ-
ical (see Figure 4) values of Tef f and log g: (4000
K and 1.4 dex), (4400 K and 2.6 dex), and (5000
K and 3.0 dex). The [Fe/H] took the values -2,
-1, and 0 dex for each case. The resulting ranges
of the estimated parameters (1-sigma level) are
shown by the solid curves in Figure 5 about the
mean values (designated by the diamonds). The
dashed lines represent the initial true parameters
(log g or [Fe/H]) of the spectra. The figure shows
that we obtain an accuracy in log g of the order of
0.2 dex. A systematic difference in log g of +0.1
dex is found and is primarily the result of the dis-
placement of the microturbulent velocity ξ from
the adopted 2 km s−1. The metallicity is defined
with 0.1 dex accuracy and a possible displacement
of all estimated values (again due to the unknown
ξ) is -0.05 dex. To check if these results depend
on the number of Monte Carlo simulations, we also
evaluated some models based on sets of 400 dete-
riorated spectra instead of 100. No obvious differ-
ences can be seen compared with the case of 100
simulations.
We also extend the considered S/N ratio up
to 100. High values of S/N do not improve the
inferred parameters since the ξ uncertainty in-
troduces the most significant error. The next
strongest factor introducing a systematic devia-
tion between the "real" and estimated stellar pa-
rameters is the RV error. It is seen that starting
with S/N = 10 and better one can estimate sur-
face gravity and metallicity from our spectra with
quite good accuracy. Most of our spectra in the
northern sample satisfy this criterion. Note that
the accuracy of the RVs also depend on S/N . This
can be a reason for more significant errors in log g
and [Fe/H] obtained from our spectra.
3.4. Stellar Parameters Derived from a
Classical Spectroscopic Analysis
Some of the higher S/N spectra from GGSS
stars studied here are suitable for a detailed abun-
dance analysis from which stellar parameters and
iron abundances can be obtained via measure-
ments of individual equivalent widths from a se-
lected sample of Fe I and Fe II lines. This is a dif-
ferent technique from the straightforward match-
6
ing of observed and synthetic spectra used in Sec-
tion 3.2, which we will refer to as the "automated
spectroscopic" method. When a strictly spectro-
scopic method (using a line-by-line analysis) is
adopted, the effective temperatures can be ob-
tained by forcing a zero slope in the relation be-
tween Fe I abundances with line excitation poten-
tials; surface gravities are obtained from the agree-
ment between the abundances from Fe I and Fe II
lines. Another parameter that can be adjusted
at the same time is the microturbulence velocity,
which is tuned so that the Fe I abundances are
independent of the equivalent widths.
This method can provide a consistency check
on the metallicities and log g derived with the au-
tomated method presented in Section 3.2. To test
the degree of consistency, a small sub-sample of
10 target stars, with larger than average S/N val-
ues, were selected as candidates for the detailed
spectroscopic analysis. Our approach consisted of
adopting the same effective temperatures from the
Alonso et al. (1999) photometric calibration for
the target stars and to use a sample of Fe I and
Fe II lines that were tested to produce good re-
sults for the Sun, as well as the well-studied red-
giant Arcturus. The sampled Fe lines and their
atomic parameters are listed in Table 2: for each
transition the excitation potential and gf-value are
listed. The equivalent-width measurements are
presented in Table 3. The same spectrum analysis
code MOOG was used in the abundance compu-
tations based on the individual equivalent-width
measurements. The adopted effective tempera-
tures, and derived surface gravities, microturbu-
lent velocities, and iron abundances are presented
in Table 4. A comparison of surface gravities de-
rived using the Fe I and Fe II lines with those
obtained from the automated method finds a sys-
tematic mean offset of [log g(automated) - log g(Fe
I/Fe II)]= +0.4±0.4 dex. As the Fe II lines are
quite sensitive to stellar surface gravity, this sys-
tematic offset may suggest that the surface gravi-
ties derived from our automated method are some-
what overestimated.
The derived Fe abundances from this detailed
analysis are intercompared in Figure 6 with the
metallicities obtained from Washington + DDO51
photometry (upper panel) and automated spec-
troscopy (middle panel). The metallicity distri-
butions shown in the middle and lower panels of
the figure are in generally good agreement, with
the same standard deviation of 0.18 dex, but an
an offset of 0.18 dex in the average [Fe/H]. Such
an offset is within the quoted uncertainties in the
metallicities obtained from the automated method
(of roughly 0.2 to 0.3 dex). Moreover, because
the automated method does not measure a true
Fe abundance, but an overall general metallicity
affected by a mixture of elements, a mild offset
between the two methods is not unexpected.
An important aspect of this additional verifica-
tion of the results obtained with the automated
method is a check on the derived surface gravities
using Fe I and Fe II lines directly, via ionization
equilibrium. As discussed previously, the GGSS
targets were selected from Washington photome-
try two-color diagrams to be giant stars and not
dwarfs. The values of log g derived for this sub-
sample of stars, as can be seen in Table 4, confirms
their evolved status as red giants. Both the direct
analysis of Fe I and Fe II ionization equilibrium,
as well as the automated direct comparison with
model spectra, indicate that the Washington +
DDO51 two-color diagram is an effective method
for identifying red giants.
4. RV Stability Versus Stellar Parameters
We now focus on the 148 stars of the north-
ern sample whose RVs were measured two or more
times with large epoch differences (a time between
successive observations of the same star not less
than 4 months). From these data we estimate the
RV dispersion which we designate as σ and utilize
this as a measure of variability of the RV. Note,
this is not the statistical standard deviation widely
designated by σ; the present σ is calculated from
multiple RV measures after weighting each by the
inverse square of the accuracy of each RV measure
(wj ): RVm = Σ RVj · wj / Σ wj, and
σ = qΣ wj · (RVj − RVm)2 / Σ wj.
(1)
Here wj is the weight. If the accuracy of the corre-
sponding RV was better than 50 m s−1, we assume
it equals 50 m s−1 because of the wavelength cal-
ibration uncertainty (see §3.2).
The derived values of σ are shown in Table 5.
Figure 7 (upper panel) shows the distribution of
the σ values (in km s−1). The largest number of
sources occur in the first bin reveals the accuracy
7
of our estimates and confirms that its value is of
the order of 50 m s−1, as it follows from the accu-
racy of the wavelength calibration (section §2.2).
We next incorporate the repeated RV data ob-
tained with CORALIE for the southern GGSS
subsample (Arenas et al. 2002) in Figure 7 (lower
panel) with 341 stars in this sample (see Table 6).
Variation of the radial velocity σ for the southern
stars observed two and more times is estimated
by Equation (1). A gaussian curve correspond-
ing to 100 m s−1 sigma is shown in Figure 7 by
dashed line. Both southern and northern subsam-
ples exhibit similar behavior in their respective
RV-variability distributions.
We found no large, apparent systematic differ-
ence in the stellar parameters between the RV-
stable (low σ) and unstable (high σ) fractions of
the sample. However, for stars with σ < 100 m
s−1 the ranges of the stellar parameters appear to
be narrower than for red giants with σ > 100 m
s−1. Thus, the fraction of RV-stable stars is low
among very cool (Tef f < 4000 K) as well as metal-
poor stars with [Fe/H] < -1.0, in agreement with
previous conclusions by Jorissen et al. (1997) and
Carney et al. (2003).
The fraction of RV-stable and unstable stars
can be estimated from the distributions of Tef f ,
log g, [Fe/H], and absolute magnitude MV pre-
sented in the left-hand panels of Figure 8; here the
distributions of stars with σ < 100 m s−1 (dashed
lines) are shown in comparison with the distribu-
tions for all stars (dotted lines). An assessment of
MV is made with Tef f , log g, and an assumption of
fixed mass for the red giants (we assume it equals
to 0.9 M⊙; variation of the mass introduces in-
significant scatter in MV ). The right-hand panels
of Figure 8 show the fraction of stable stars in each
bin of the histograms shown on the left. As seen in
the figure, the fraction of stable stars has a peak
in the range 4300-4700 K (secondary peaks at the
wings of the distribution are caused by small num-
ber statistics). Note that this range corresponds to
the extinction-corrected (J − K)0 values from 0.59
to 0.73 mag, and these values do not depend on
the choice of color-temperature calibration. Fig-
ure 2 shows that the bulk of the giant stars selected
from the GGSS Washington+DDO51 photometry
survey occupies this range of temperatures, show-
ing that the GGSS happens to be optimized for
selecting stars in the most RV-stable temperature
8
range.
A similar peak takes place at log g = 2.3 -- 3.2 in
the surface gravity distribution. We also find that
the largest fraction of RV-stable stars happens for
stars of slightly subsolar metallicity (-0.5 ≤ [Fe/H]
≤ -0.1), and then the fraction of stable stars ap-
pears to drop precipitously near solar metallicity.
Note, however, that the latter conclusion needs for
an additional check since the most metal rich bin
is only sampled by 6 stars.
An analog of Figure 8 is drawn for the southern
subsample in Figure 9. Here we use the "photo-
metric" [Fe/H] derived by Rhee et al. (2001). The
surface gravity is established by way of interpo-
lation of the isochrones published by Bergbusch
& Vandenberg (1992). The histograms for Tef f
defined by the same method for both subsamples
of northern and southern stars, are qualitatively
similar in Figures 8 and 9. Though the other pa-
rameters are estimated by different ways, the cor-
responding histograms for [Fe/H], log g, and MV
are qualitatively rather similar.
In order to better constrain what types of stars
may be most RV-variable we again focus on the
McDonald data and the stellar parameters derived
from these spectra. In Figure 10 the values of log g
are plotted versus the effective temperatures: this
is a spectroscopic version of an HR-diagram. The
top panel contains those stars with RV variability
of less than 100 m s−1, while the bottom panel are
the definite RV variables having σ > 100 m s−1.
Superimposed on these stellar points are isochrone
curves from Girardi et al. (2000); two metallic-
ity isochrones are shown (the solid curves have
[Fe/H]=0.0 while the dashed curves have [Fe/H]=-
0.7) with two ages for each metallicity (3.5 and 8.9
Gyr). Plotted this way, the initial RV-results indi-
cate that those stars exhibiting variability of less
than ∼100 m s−1 tend to be more concentrated in
the log g − Tef f plane, near the He-core burning
clump for near-solar metallicities. The stars tend-
ing to show RV-variability fall all along either well-
defined first-ascent RGB, or AGB, both at some-
what low metallicities. Given uncertainties in both
the models and the gravities and temperatures de-
rived by us (which almost certainly also carry some
systematic differences), the concentration of pos-
sibly RV-stable red giants near Tef f ∼4500K and
log g ∼2.5-3.0 coincides quite closely to the red gi-
ant clump for stars having metallicities near solar,
or slightly lower. Note that real log g may have
systematically lower values than those shown in
Figure 10 (see also §3.4).
The possible association of quasi-RV stability
with the red giant clump can be investigated fur-
ther by including the southern sample, although
we do not have spectroscopically derived gravities
for these stars. For a relatively old population,
however, the effective temperature of the clump
does not vary much if the metallicity does not
vary by much: this can been seen by inspecting
the isochrones in Figure 10. The indications from
the McDonald results are that the RV-stable red
giants are only mildly metal poor with a relatively
small dispersion in metallicity. For the 36 Mc-
Donald RV-stable stars having Teff =4500±200K
(corresponding to the observed "clump" of points
in Figure 10), the mean metallicity is [Fe/H]=-
0.5±0.2. The combined northern and south-
ern samples are plotted as effective temperature
histograms in Figure 11, where the sample has
been segregated using σ. There is a clearly de-
fined clump of potentially RV-stable stars with
Teff ∼4500K that probably corresponds to core-He
burning red giants with mild metal deficiencies.
The stars with σ >100 m s−1, shown in the bot-
tom panel, are significantly more dispersed in Teff ,
with almost equal numbers over a range in effec-
tive temperature; our suggestion is that these are
primarily first-ascent giants or AGB stars, as well
as intrinsically RV-stable stars (such as clump gi-
ants) that have companions. The fraction of stars
whose RV-variability is caused by components is
not known from our observations. Demonstrat-
ing the presence of companions will require much
more repeated observations.
Preliminary results from monitoring of photo-
metric variability of a subsample of GGSS stars
by Bizyaev et al. (2004) reveal no correlation be-
tween RV- and brightness variations. However,
more precise and simultaneous photometric and
spectroscopic observations are needed to shed light
on the nature of the RV variations in red giants.
4.1. Strategies for Selecting SIM Astro-
metric Grid Stars
Our analysis gives insight into the optimal se-
lection of giant stars for the SIM Astrometric Grid
if time and resources are limited (as indeed they
are) for ground-based vetting of large numbers of
9
stars for the most RV-stable specimens.
As can be seen in Figure 11, a higher fraction
of stars with σ < 100 m s−1 can be found with
4300 K < Tef f < 4700 K. This is a simple,
straightforward selection criterion resulting in an
initial success rate of 50% in finding RV-stable red
giants. As a point of reference, the entire northern
sample contains 39% stable stars and the southern
sample contains 32%.
If high-resolution spectra
are also used to derive surface gravities and metal-
licities, the additional constraints of -0.5 < [Fe/H]
< -0.1, and 2.3 < log g < 3.2 lead to a 81%
and 40% success rate for the northern and south-
ern samples, respectively, yet significantly narrows
the sample. Note that the northern sample re-
stricted simultaneously by all stellar parameters
reveals the fraction of RV-stable stars that is close
to findings by Frink et al. (2001).
More attention to exploring the solar and
higher metallicity ranges is warranted because if
the relatively low RV-stable fraction we find at
these solar-like metallicities is borne out it would
have serious ramifications for the currently se-
lected SIM Astrometric Grid sample, which is
dominated by giant stars randomly selected from
the Tycho catalogue. Spectroscopic metallicities
measured for nearly 500 Tycho giants by J. Crane
(in preparation) in the preferred magnitude range
(V < 11) for the SIM Grid stars indicates a me-
dian metallicity for this sample of [Fe/H] = -0.1,
and perhaps larger (i.e., 0.0 dex) if the arguments
for the need of a +0.1 dex correction to the local
metallicity scale are applied (Reid 2002; Haywood
2002). Thus, a significant fraction of the Tycho
sample is at metallicities where a large fraction of
stars may have troublesome atmospheric jitter.
5. Conclusions
We are carrying out high-resolution spectro-
scopic observations of an all-sky sample of giants
which was pre-selected based on the GGSS pro-
gram of Washington photometry. The GGSS was
intended to identify stars with the highest like-
lihood of astrometric stability -- namely, subso-
lar metallicity red giants. The main purpose of
the present study is to make an initial assess-
ment of the RV-variability properties of GGSS
stars as a function of their atmospheric charac-
teristics. The RV variation was estimated for 489
stars with spectroscopy taken on both Northern
and Southern telescopes. Basic stellar parameters
were estimated for the northern subsample from
high-resolution spectroscopy.
A surprisingly high fraction of investigated gi-
ant stars have unstable radial velocities at 100 m
s−1 level: about 2/3 of our sample. Both of the
samples, northern and southern, although having
been observed independently and with different
instruments and techniques, show similar distri-
butions of RV variability. Although a number of
obvious spectroscopic binaries are included in this
sample, much of the low-amplitude RV variabil-
ity is probably due to atmospheric motions and
temperature inhomogeneities.
A higher fraction of stars with σ < 100 m s−1
can be found among the objects with 4300 K <
Tef f < 4700 K, which corresponds to (J − K)0 =
0.59 -- 0.73. If we incorporate spectroscopic metal-
licity and surface gravity information as well, and
select only stars with -0.5 < [Fe/H] < -0.1 and
2.3 < log g < 3.2, it will help to identify RV-
stable candidates more effectively, but narrows the
sample of potential candidates substantially. From
the point of view of minimizing expensive obser-
vations, the optimal way of preliminarily select-
ing RV-stable stars is to use the effective temper-
ature, or more exactly, the calibration indepen-
dent range of NIR colors as 0.59 < (J − K)0 <
0.73. This range encompasses 44% of stars with
σ < 100 m s−1 in both our southern and northern
samples; 38.5% of the initial GGSS sample of can-
didate Astrometric Grid stars (1710 of 4440) lies
in this range. The result obtained here will aid
in the continuing efforts to define the astrometric
reference grid for SIM.
The GGSS follow-up observations presented
here have been supported by the JPL and NASA
via grant 99-04-OSS-058. The photometric por-
tion of the GGSS was funded by NASA/JPL
grants 1201670 and 1222563 to SRM and RJP, and
with substantial contributions of additional fund-
ing from a David and Lucile Packard Foundation
Fellowship to SRM. VVS acknowledges support
for part of this work from NASA via grant NAG5-
13175. We are grateful to the staff of McDonald
observatory for their help and support of both
photometric and high resolution spectroscopic ob-
servations, and from the Carnegie Observatories
for similarly generous contributions for photomet-
ric and medium resolution spectroscopic observa-
tions of the GGSS. 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 IPAC/Caltech, funded
by the NASA and NSF.
REFERENCES
Adelman, S., Baltic Astronomy 2001, 10, 593
Alonso, A., Arribas, S., Martinez-Roger, C. 1999,
A&AS, 140, 261
Arenas, J., Geisler, D., Majewski, S., Smith, V. et
al. 2002 AAS, 201, 115.05
Baranne, A., Queloz, D., Mayor, M. et al. 1996,
A&AS, 119, 373
Bergbusch, P., Vandenberg, D. 1992, ApJS, 81,
163
Bessell, M., Castelli, F., Plez, B. 1998, A&A, 333,
231
Bizyaev, D., Smith, V. V., Arenas, J., Geisler, D.
2004, AAS, 205, 91.07
Carney, B., Latham, D., Stefanik, R. et al. 2003,
AJ, 125, 293
Cote, P., Pryor, C., McClure, R. D., Fletcher, J.
M., & Hesser, J. E. 1996, AJ, 112, 574
Fischer, D., Valenti, J. 2005, ApJ, 622, 1102
Frink, S., Quirrenbach, A., Fischer, D., Roser, S.,
& Schilbach, E. 2001, PASP, 113, 173
Girardi, L., Mermilliod, J.-C., & Carraro, G. 2000,
A&AS, 141, 371
Griffin, R., Lynas-Gray, A. 1999, AJ, 117, 2998
Hatzes, A., Cochran, W. ApJ1994, 432, 763
Haywood, M. 2002 ,MNRAS, 337, 151
Hinkle, K., Wallace, L., Valenti, J., & Harmer, D.
Visible and Near Infrared Atlas of the Arcturus
Spectrum, 3727-9300 A 2000, (San Francisco:
ASP) ISBN: 1-58381-037-4
Jorissen, A., Mowlavi, N., Sterken, C., Manfroid,
J. A&A1997, 324, 578
10
Kurucz, R. L. 1993, CD-ROM, Cambridge, MA:
Smithsonian Astrophysical Observatory, De-
cember 4, 1993
Majewski, S. R., Ostheimer, J. C., Kunkel, W. E.,
& Patterson, R. J. 2000, AJ, 120, 2550
Nidever, D. L., Marcy, G.W., Butler, R.P., Fis-
cher, D.A., Vogt, S.S. 2002, ApJS, 141, 503
Palmer, B. A., Engleman, R., Jr. 1983, Atlas of
the Thorium Spectrum, Los Alamos National
Laboratory, Los Alamos, New Mexico
Patterson, R. J., Majewski, S. R., Slesnick, C. L.,
et al. Small Telescope Astronomy on Global
Scales, ASP Conf. Ser. v.246,
IAU Collo-
quium 183. Ed. by Bohdan Paczynski, Wen-
Ping Chen, and Claudia Lemme. San Francisco:
Astronomical Society of the Pacific, 2001, p.65
Pryor, C. P., Latham, D. W., & Hazen, M. L.
1988, AJ, 96, 123
Reid, I. N. 2002, PASP, 114, 306
Rhee, J., Slesnick, C., Crane, J., Polak, A. et al.
2001 AAS 198, 62.03
Sneden, C. 1973 Ph.D. thesis, Univ. of Texas,
Austin
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
11
Stellar Parameters for the entire McDonald Sample of 434 GGSS stars.
Table 1
Name
RA (J2000.0)
(HH:MM:SS)
DEC (J2000.0)
(DD:MM:SS)
G2358+00.31
G2358+00.92
G0001+00.94
G0011+05.87
G0011+16.75
...
00:00:51.96
00:01:36.38
00:04:20.08
00:14:18.85
00:14:26.65
00:22:16.8
00:14:09.9
00:27:12.2
05:57:37.5
17:16:03.0
Tef f
K
4290
4399
4324
4440
4594
[Fe/H]ph
(dex)
Dph
(Kpc)
log gsp
(dex)
∆ log gsp
(dex)
[Fe/H]sp
(dex)
∆[Fe/H]sp
(dex)
-0.3
-0.9
-0.7
-0.5
-0.8
1.3
1.1
3.0
0.7
1.8
4.87
2.24
2.67
2.13
1.47
0.19
0.13
0.41
0.19
0.38
-2.80
-0.42
-0.40
-0.47
-0.67
0.05
0.05
0.16
0.09
0.19
Note. -- Name of star, its RA and DEC, effective temperature, photometric metallicity [Fe/H], photometric distance, spectroscopic surface
gravity log g and its accuracy, and spectroscopic metallicity and its accuracy.
The full version of the Table is presented in electronic form.
12
Table 2
The Fe I & Fe II Lines
λ(A)
Species χ(eV)
log gf
5307.361
5322.041
5497.516
5501.464
5522.447
5536.583
5539.284
5549.948
5559.638
5560.207
5577.031
5579.335
5607.664
5608.974
5611.361
5618.631
5619.224
5636.696
5638.262
5661.012
5677.684
5679.025
5696.103
5698.023
5705.466
5717.835
5724.454
5759.261
5760.345
5784.657
5793.913
5806.717
5807.782
5809.217
5811.917
5814.805
5837.700
5838.370
5844.917
5845.266
5849.682
1.608
2.279
1.011
0.958
4.209
2.832
3.642
3.695
4.988
4.435
5.033
4.231
4.154
4.209
3.635
4.209
3.695
3.640
4.220
4.580
4.103
4.652
4.549
3.640
4.301
4.284
4.284
4.652
3.642
3.397
4.220
4.608
3.292
3.884
4.143
4.283
4.294
3.943
4.154
5.033
3.695
-2.970
-2.840
-2.840
-2.957
-1.400
-3.812
-2.660
-2.904
-1.829
-1.040
-1.551
-2.406
-2.258
-2.402
-2.993
-1.260
-3.256
-2.608
-0.720
-2.432
-2.694
-0.770
-1.997
-2.689
-1.360
-0.980
-2.627
-2.073
-2.490
-2.673
-1.697
-0.900
-3.404
-1.690
-2.427
-1.820
-2.337
-2.337
-2.940
-1.818
-2.993
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
13
Table 2 -- Continued
λ(A)
Species χ(eV)
log gf
5853.149
5856.083
5861.107
5000.743
5018.440
5132.669
5234.625
5264.812
5284.098
5325.559
5414.046
5425.247
5534.847
Fe I
Fe I
Fe I
Fe II
Fe II
Fe II
Fe II
Fe II
Fe II
Fe II
Fe II
Fe II
Fe II
1.485
4.294
4.283
2.778
2.891
2.807
3.221
3.230
2.891
3.221
3.221
3.199
3.244
-5.268
-1.640
-2.452
-4.745
-1.213
-4.000
-2.240
-3.188
-3.010
-3.170
-3.620
-3.210
-2.770
Note. -- Adopted excitation poten-
tials and gf-values for Fe I and Fe II
lines
14
Table 3
Equivalent-Width Measurements
λ
#1 #2 #3 #4 #5 #6 #7 #8 #9 #10
5307.361
5322.041
5497.516
5501.464
5522.447
5536.583
5539.284
5549.948
5559.638
5560.207
5577.031
5579.335
5607.664
5608.974
5611.361
5618.631
5619.224
5636.696
5638.262
5661.012
5677.684
5679.025
5696.103
5698.023
5705.466
5717.835
5724.454
5759.261
5760.345
5784.657
5793.913
5806.717
5807.782
5809.217
5811.917
5814.805
5837.700
5838.370
5844.917
5845.266
5849.682
132
92
...
...
...
35
42
32
18
63
...
...
...
25
31
70
21
40
98
...
20
66
27
...
50
76
19
...
44
52
51
60
27
70
...
36
...
...
...
...
...
109
84
183
...
50
25
...
15
...
...
...
14
...
...
...
58
11
...
78
...
...
54
17
28
48
69
...
...
...
36
38
55
...
54
15
34
13
25
...
...
14
185
140
...
...
...
...
54
44
19
...
22
34
44
35
37
91
195
65
126
15
24
...
...
50
88
105
19
20
65
77
71
90
41
98
35
59
...
...
...
...
31
117
86
172
166
47
17
23
...
...
57
...
10
...
...
14
56
6
24
78
...
...
53
...
...
44
66
...
...
29
34
39
49
16
53
...
...
12
21
...
...
...
118
85
...
...
...
...
32
...
...
60
...
...
25
...
19
62
10
...
89
...
15
63
...
31
54
67
...
13
41
42
46
56
23
56
...
37
16
32
...
...
17
154
100
...
...
...
...
45
...
15
68
...
25
35
31
37
71
21
...
92
13
24
72
...
47
65
81
...
50
23
57
54
66
39
70
29
48
28
45
16
14
30
165
118
...
...
...
...
44
31
11
68
...
23
33
...
27
76
15
40
98
...
20
74
...
...
...
82
...
15
...
60
57
66
30
79
25
42
21
...
...
...
26
116
819
...
...
...
...
28
21
...
56
15
16
...
...
...
58
...
32
77
...
...
51
18
...
47
...
14
12
38
44
42
...
17
...
16
...
...
...
8
9
19
135
102
...
...
...
...
...
...
...
62
...
...
27
13
19
64
...
...
96
...
...
66
18
30
52
72
9
11
36
41
...
64
...
70
...
33
...
31
8
10
14
154
114
...
...
73
34
45
...
...
72
...
...
...
24
...
78
15
48
99
...
17
74
22
39
63
86
...
...
53
54
55
68
26
76
...
...
...
...
...
10
20
15
Table 3 -- Continued
λ
#1 #2 #3 #4 #5 #6 #7 #8 #9 #10
5853.149
5856.083
5861.107
5000.743
5018.440
5132.669
5234.625
5264.812
5284.098
5325.559
5414.046
5425.247
5534.847
...
51
...
...
...
39
87
45
64
...
30
46
...
32
...
...
...
175
...
71
43
63
42
30
41
60
67
70
...
...
201
37
100
55
82
54
30
52
70
22
...
...
...
...
41
92
54
76
56
37
57
73
32
42
...
...
...
35
72
44
...
47
34
46
59
55
51
...
...
222
49
100
57
...
57
...
61
74
58
53
22.7
...
...
45
86
56
...
56
42
51
70
57
56
...
...
186
...
81
44
67
45
30
44
...
...
39
...
...
174
...
72
38
62
...
29
38
...
27
...
...
...
161
25
786
42
...
36
20
38
...
Note. -- Equivalent widths of Fe I and Fe II lines (given in mA) for 10 stars. Stellar
identifications are provided in Table 4.
Stellar Parameters and Metallicities derived from classical spectroscopic analysis for
10 selected stars
Table 4
Star
ID
G0858+00.192 #1
G0901+00.175 #2
#3
G1007+00.5
G1018-05.46
#4
G1030+00.15 #5
G1042-05.69
#6
G1053+00.4
#7
G1113+00.20 #8
G1147-05.31
#9
G1147-05.49 #10
Teff
(K)
4630
4760
4490
4670
4780
4370
4400
4240
4600
4720
log g
(dex)
ξ
(km/s)
[Fe/H]
(dex)
2.2
2.6
2.5
1.6
2.6
1.1
1.2
1.6
2.1
3.0
1.1
0.9
1.8
1.2
0.9
1.4
0.9
1.4
0.6
1.3
-0.31 ±0.07
-0.48 ±0.06
-0.12 ±0.05
-0.77 ±0.05
-0.34 ±0.06
-0.63 ±0.06
-0.35 ±0.07
-0.56 ±0.06
-0.48 ±0.07
-0.41 ±0.08
Note. -- Name of star, its ID in Table 3, effective tempera-
ture, surface gravity, microturbulent velocity, and [Fe/H].
16
RV-variability for the 148 McDonald Repeated Stars
Table 5
Name
G0118-05.38
G0131+00.86
G0142+05.56
G0142+05.70
G0151+00.72
...
M
(mag)
11.97
11.55
10.91
12.27
11.29
Teff
(K)
4814
4707
3959
4196
4334
Dph
(Kpc)
[Fe/H]ph
(dex)
log gsp ∆ log gsp
(dex)
(dex)
[Fe/H]sp
(dex)
∆ [Fe/H]sp
(dex)
σ
(km/s)
0.9
0.9
2.2
2.3
1.7
-1.0
-1.2
-0.9
-0.6
-0.6
3.13
2.34
1.20
1.93
2.34
0.13
0.17
0.65
0.25
0.06
0.00
-0.80
-0.60
-0.67
-0.07
0.05
0.14
0.28
0.25
0.22
0.165
0.919
0.489
0.287
0.287
N
2
2
2
2
3
Note. -- Name of star, Washington M magnitude (which is approximately Johnson's V ), effective temperature, photometric
distance, photometric metallicity [Fe/H], spectroscopic surface gravity log g and its accuracy, spectroscopic metallicity [Fe/H]
and its accuracy, RV variability, and number of RV-observations for the northern sample objects.
The full version of the Table is presented in electronic form.
RV-variability for the 341 CORALIE Repeat Observation Stars
Table 6
Name
M
(mag)
Tef f
(K)
Dph
(Kpc)
[Fe/H]ph
(dex)
σ
N
(km/s)
G0000-56.81
G0012-28.38
G0016-39.1207
G0016-39.2075
G0016-39.3290
...
11.92
11.71
11.46
11.93
11.41
4066
3891
4418
4580
4753
3.5
3.8
4.3
1.7
0.8
-1.1
-1.0
-2.0
-1.0
-0.7
0.004
0.182
1.104
0.635
2.060
3
3
7
2
4
Note. -- Name of star, Washington M magnitude, effective tempera-
ture, photometric distance, photometric metallicity [Fe/H], RV variabil-
ity, and number of RV observations for objects of the southern sample.
The full version of the Table is presented in electronic form.
17
Fig.
1. -- An example of the order-to-order
radial velocity differences for the program star
G1113+00.20. The wavelength corresponds to the
middle of each respective order.
Fig. 2. -- Distribution of Tef f for the general sam-
ple of 4440 GGSS candidates.
18
Fig.
3. -- The comparison of metallicities es-
timated from high-resolution spectroscopy (see
text) and "photometric" [Fe/H] (Rhee et al. 2001).
The dashed line shows the "robust" least absolute
deviation fit to the data. The values of [Fe/H]
for most of the stars with intermediate metallicity
approximately follow a linear relation with disper-
sion of order 0.4 dex. The spectroscopic [Fe/H] is
systematically higher than the photometric one by
about 0.16 dex.
Fig. 4. -- The relation between surface gravity and
effective temperature for 434 stars from our north-
ern sample.
Fig. 5. -- The 1-sigma level scatter of estimated
stellar atmospheric parameters (solid curves) as
a function of S/N . The mean values of log g
and [Fe/H] found for one hundred deteriorated
model spectra are designated by the diamonds.
The dashed lines represents the initial (true) pa-
rameters (log g or [Fe/H]) of the model spec-
tra. The results of simulations are shown for
models typical of the GGSS red giants: Tef f =
4000,
log g = 2.6, and
log g = 3.0, and for metallicities
Tef f = 5000,
[Fe/H]=0, -1, and -2.
log g = 1.4, Tef f = 4400,
19
Fig. 6. -- Metallicities for 10 GGSS stars measured
three different ways: using Washington photome-
try, the automated method, and a detailed spec-
troscopic analysis. The three methods show gen-
eral agreement.
Fig. 7. -- Distribution of the variation of radial
velocity, σ (in km s−1) for the northern (upper
panel) and southern (lower panel) subsamples (148
and 341 stars, respectively) with two or more es-
timates of RV. The dashed line shows a gaussian
distribution with 100 m s−1 sigma).
Fig. 8. -- Left panels: histograms of distribution
of Tef f , log g, [Fe/H], and MV for the stars whose
σ < 100 m s−1 (dashed line) in comparison with
the whole sample (dotted line). The right side
panels show the corresponding fraction of stable
stars in bins of the histogram.
20
Fig. 9. -- The same as in Figure 9, but for the
southern subsample.
0
2
4
0
2
4
6000
5000
4000
3000
Fig. 10. -- The spectroscopically derived surface
gravity, log g, plotted versus Teff for the McDon-
ald northern sample: The top panel shows stars
with σ < 100 m s−1 while the bottom panel shows
those with RV-variability greater than 100 m s−1.
The RV-stable stars are more concentrated in the
log g -- Tef f plane than the variables. The con-
tinuous curves are isochrones from Girardi et al.
(2000), with the dashed curves having [Fe/H]=-
0.7 and the solid curves [Fe/H]=0.0. The two
isochrones for each metallicity have different ages
of 3.5 Gyr and 8.9 Gyr, respectively. The concen-
trated "clump" of RV-stable stars may correspond
to the core-He burning red giant phase of stellar
evolution.
21
40
30
20
10
0
40
30
20
10
0
4000
4500
5000
Fig. 11. -- Effective temperature histograms seg-
regated as the RV-stable stars (top panel) and
RV-variable stars (bottom panel) for the combined
northern and southern data sets. The RV-stable
stars are sharply peaked near Tef f =4500±200K,
while the RV-variable stars are more spread out
in temperature. This provides additional support
for the suggestion that the core-He burning stars,
with only mild metal deficiencies, may be intrinsi-
cally more stable in radial velocity.
22
|
astro-ph/0703390 | 1 | 0703 | 2007-03-15T11:17:47 | The circumstellar envelope of IRC+10216 from milli-arcsecond to arcmin scales | [
"astro-ph"
] | Aims.Analysis of the innermost regions of the carbon-rich star IRC+10216 and of the outer layers of its circumstellar envelope have been performed in order to constrain its mass-loss history. Methods: .We analyzed the high dynamic range of near-infrared adaptive optics and the deep V-band images of the circumstellar envelope of IRC+10216 using high angular resolution, collected with the VLT/NACO and FORS1 instruments. Results: .From the near-infrared observations, we present maps of the sub-arcsecond structures, or clumps, in the innermost regions. The morphology of these clumps is found to strongly vary from J- to L-band. Their relative motion appears to be more complex than proposed in earlier works: they can be weakly accelerated, have a constant velocity, or even be motionless with respect to one another. From V-band imaging, we present a high spatial resolution map of the shell distribution in the outer layers of IRC+10216. Shells are resolved well up to a distance of about 90'' to the core of the nebula and most of them appear to be composed of thinner elongated shells. Finally, by combining the NACO and FORS1 images, a global view is present to show both the extended layers and the bipolar core of the nebula together with the real size of the inner clumps. Conclusions: .This study confirms the rather complex nature of the IRC+10216 circumstellar environment. In particular, the coexistence at different spatial scales of structures with very different morphologies (clumps, bipolarity, and almost spherical external layers) is very puzzling. This confirms that the formation of AGB winds is far more complex than usually assumed in current models. | astro-ph | astro-ph | Astronomy&Astrophysicsmanuscript no. Leaoirc10216referee
July 26, 2021
c(cid:13) ESO 2021
7
0
0
2
r
a
M
5
1
1
v
0
9
3
3
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
The circumstellar envelope of IRC+10216 from milli-arcsecond to
arcmin scales ⋆
I.C. Leao1,2, P. de Laverny1, D. M´ekarnia1, J.R. De Medeiros2, and B. Vandame3
1 Observatoire de la Cote d'Azur, Dpt Cassiop´ee, CNRS - UMR 6202, BP 4229, 06304 Nice Cedex 4, France
e-mail: [leao;laverny;mekarnia]@obs-nice.fr
2 Departamento de F´ısica, Universidade Federal do Rio Grande do Norte, 59072-970 Natal, RN, Brazil
3 European Southern Observatory, Karl-Schwarzschild-Str. 2, D -- 85748 Garching b. Munchen, Germany
Received 23 November 2005 / Accepted 15 January 2006
ABSTRACT
Aims. Analysis of the innermost regions of the carbon-rich star IRC+10216 and of the outer layers of its circumstellar envelope have been
performed in order to constrain its mass-loss history.
Methods. High dynamic range near infrared adaptive optics and high angular resolution deep V-band images of its circumstellar envelope
collected with VLT/NACO and VLT/FORS1 instruments have been analyzed.
Results. Maps of the sub-arcsecond structures, or clumps, in the innermost regions are derived from the near-infrared observations. The
morphology of these clumps is found to strongly vary from J- to L-band. Their relative motion appears to be more complex than proposed
in earlier works: they can be weakly accelerated, have a constant velocity, or even be motionless with respect to one another. From V-band
imaging, a high spatial resolution map of the shell distribution in the outer layers of IRC+10216 is presented. Shells are well resolved up to a
distance of about 90′′ to the core of the nebula and most of them appear to be composed of thinner elongated shells. Finally, by combining the
NACO and FORS1 images, a global view, showing both the extended layers and the bipolar core of the nebula together with the real size of the
inner clumps is presented.
Conclusions. This study confirms the rather complex nature of the IRC+10216 circumstellar environment. In particular, the coexistence at
different spatial scales of structures with very different morphologies (clumps, bipolarity and almost spherical external layers) is very puzzling.
This confirms that the formation of AGB winds is far more complex than usually assumed in current models.
Key words. stars: AGB and post-AGB -- stars: variables: general -- stars: individual: IRC+10216 -- stars: mass-loss -- stars: circumstellar matter
-- techniques: high angular resolution
1. Introduction
Low- and intermediate-mass stars lose a large amount of their
initial mass when they evolve along the Asymptotic Giant
Branch (AGB) and beyond. During these mass-loss events, a
huge circumstellar envelope (CSE) is formed. IRC+10216 is
the best-known example of such evolved stars with an optically
thick CSE. Indeed, its envelope almost completely absorbs the
central stellar photons in visible light and at shorter wave-
lengths. This circumstellar environment has therefore been
mostly studied in the infrared and millimeter domains, spec-
tral regions where the envelope radiates itself and scatters the
stellar light.
At very small scales (arcsec and below), a detailed picture
of the IRC+10216 central regions has already been presented
⋆ Based on observations collected with the VLT/Antu and Yepun
telescopes (Paranal Observatory, ESO, Chile) using the FORS1 and
NACO instruments (programs ID 63.I-0177A, 70.C-0565A, 70.C-
0271B, 70.D-0271B).
by several groups (see e.g. Haniff & Buscher 1998; Weigelt et
al. 1998, 2002; Tuthill et al. 2000, 2005). The central core ap-
pears to be composed of a series of clumps whose positions
and luminosities vary on time-scales of a few years. The com-
plexity of the structures detected has led to several hypotheses
regarding the precise location of the central star.
At much larger scales
(up to arcmin), Mauron &
Huggins (1999, 2000, MH99-00 hereafter) have shown that the
IRC+10216 CSE can also be studied in visible wavelength if
enough deep images are collected. The nebula brightness then
results from galactic ambient light scattered by its dust parti-
cles. It is detected up to very large distances to the central star
(up to about 6 000 stellar radii) and thus carries information
about the mass-loss history during the last few thousand years.
MH99-00 have also shown that this fairly round circumstel-
lar envelope is consistent with an isotropic galactic radiation
field and a spherically symmetric dust shell (see also Mauron et
al. 2003). However, on a better spatial resolution (∼ 1 arcsec),
the envelope consists of a series of discrete and nested multi-
2
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
ple shells (or arclets) whose origin is still debated. Although
IRC+10216 is the only known AGB with such shells, similar
morphology has already been detected around a dozen of plane-
tary nebulae (PN) and about six proto-planetary nebulae (PPN).
However, all these PN and PPN are bipolar, contrary to what we
observe for their progenitor (assuming that IRC+10216 CSE
properties are common for AGB stars). The cause and occur-
rence of the transition from a spherical multiple-shell CSE to a
bipolar one is crucial for the understanding of the mass-loss
phenomenon on the AGB and the evolution of the material
ejected into the interstellar environment.
the morphology of
To date, no global view of
the
IRC+10216 CSE at different scales exists. The aim of this work
is to provide such a global description by combining new high
dynamic and high spatial resolution images of its innermost re-
gions collected with adaptive optics techniques together with
new deep images of its most external layers. These observa-
tions are presented in Sect. 2. We analyze in Sect. 3 the mor-
phology of the innermost regions and their temporal variations.
Sect. 4 is devoted to the analysis of the numerous shells found
in this envelope and to some of their properties. We then dis-
cuss, in Sect. 5, the coexistence of the different morphologies
found in the CSE of IRC+10216. Finally, a conclusion is pre-
sented in Sect. 6.
2. Observations and reductions
2.1. NACOobservations
Infrared images of IRC+10216 were recovered from ESO
Science Archive Facility. They were obtained in November
2002 and March 2003, using the adaptive optics system NACO
at the ESO/VLT Yepun telescope. NACO is an association
of the adaptive optics system NAOS (Rousset 2000) and the
spectro-imager CONICA (Lenzen 2003).
We have recovered observations of IRC+10216 obtained
with the narrow-band filters NB 1.24 (centered at λc =
1.237 µm, ∆λ = 0.015 µm), NB 1.26 (λc = 1.257 µm, ∆λ =
0.014 µm), NB 1.64 (λc = 1.644 µm, ∆λ = 0.018 µm), NB 1.75
(λc = 1.748 µm, ∆λ = 0.026 µm), NB 2.17 (λc = 2.166 µm,
∆λ = 0.023 µm) and the broad-band filter L′ (λc = 3.80 µm,
∆λ = 0.62 µm). The pixel scale on CONICA was respectively
13.27 mas in the narrow-band filters and 27.15 mas in the L′ fil-
ter. Observation conditions, as well as total on-source integra-
tion time for each filter, are summarized in Tab. 1. Calibration
files (flat fields and dark exposures) and observations of the
PSF reference star HR 3550 were also recovered. The Jitter
technique was used in all observations. The box size of the L′
broad-band image was about 7′′ × 7′′, and the box sizes in the
narrow-band images varied between about 4′′ ×4′′ and 7′′ ×7′′.
As shown in Tab. 1, the seeing conditions were variable, rang-
ing between about 0.5′′ and 0.8′′. The best dynamic ranges of
the IRC+10216 final images (not deconvolved) were around
7 000, 40 000, 6 000 and 90 000 AFU (Arbitrary Flux Units)
for the J, H, K and L bands, respectively. The noise level was
found to be smaller than 30 AFU in all images. In the PSF
observations, the seeing varied between 0.5′′ and 0.6′′ and the
air-masses between about 1.3 and 1.4. The estimated FWHM
Table 1. NACO observations log of IRC+10216.
Date
Filter
(UT)
22 Nov 02 NB 1.64
NB 2.17
16 Mar 03 NB 1.26
NB 1.64
NB 2.17
18 Mar 03 NB 1.24
NB 1.75
NB 2.17
L′
On-source
exp. time
Seeing
Air-
mass
(sec)
128
120
120
70
60
210
103
200
183
(′′)
0.6
0.6
< 0.5
< 0.5
< 0.5
0.6
0.7
0.8
0.6
1.5
1.6
1.3
1.3
1.3
1.3
1.3
1.3
1.3
Dyn.
range
(AFU)
28 700
6 200
6 900
8 300
4 800
7 200
41 500
5 200
92 500
of the PSF star was around 70 mas in the J and H bands, 80 mas
in the K-band and 120 mas in the L-band.
Standard reduction procedures were applied using self-
developed routines. The raw images were sky subtracted, then
divided by the flat-field and corrected from hot pixels. In each
filter, the images were cross-correlated and aligned by sub-
pixel shifting, and then combined to produce the final images,
eliminating cosmic rays hits. Finally, they were deconvolved
with the PSF reference star. We used the Richardson-Lucy al-
gorithm (Richardson 1972; Lucy 1974). Since no PSF data for
IRC+10216 were found in the November 22, 2002 observa-
tions, we have developed for that night pseudo-PSF images,
by analyzing and comparing the other IRC+10216 observa-
tions with their corresponding PSF data. Constancy of promi-
nent features present in deconvolved images showed that the
PSF selection and the number of iterations (25 typically) for
the deconvolution process was performed carefully and con-
servatively. We have then summed the deconvolved images in
each band (see Fig. 1). The highest dynamic range J-band im-
age was obtained from the 1.24µm and 1.26µm images, which
leads to about 14 000 AFU. For the H-band, we have summed
both 1.64µm and 1.75µm narrow-band images, obtaining a dy-
namic range of about 78 000 AFU. The three 2.17µm narrow-
band images were combined to build a K-band image (∼ 16 000
AFU). Finally, the L-band image has about 92 000 AFU. These
images have thus the best dynamical range ever published (see
e.g. Tuthill et al. 2005). We note that, over the interval of about
4 months between the first and last observations studied here,
no clear variations of the positions of the structures were found.
2.2. FORS1observations
The observations were collected with the VLT-Antu telescope,
equipped with the FORS1 focal reducer. The detector is a
2048×2048 thinned 24 µm pixel Tektronix chip. The field of
view of individual images is 6.8′ × 6.8′ with a pixel size of
0.2 arcsec (see Appenzeller et al. 1998). All the exposures were
acquired in standard FORS1 service mode using a classical
Bessel V-band filter. The available data consists of eight 10-
min exposure and two 20-min exposure frames collected from
10 to 11 January, 2000, leading to a total observing time of
2 hours. The selected exposures were taken in dark time under
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
3
very good seeing conditions and photometric sky. A few other
10-min frames were indeed rejected due to their moderate see-
ing conditions. The mean airmass was 1.3. The telescope was
shifted by a few arcseconds between each individual image.
It was found that individual images reduced by the stan-
dard ESO reduction pipeline (which includes standard correc-
tions such as bias subtraction, flat fielding, etc., see Hanuschik
& Amico 2000) were of rather poor quality. We suspect this
was due to the use of a corrupted flat field. Therefore, a new re-
duction procedure was performed for all individual exposures
(removal of cosmic and aberrant pixels, flat-fielding with a spe-
cific mean sky flat for each night, etc.). All exposures taken
on the same night were then shifted and stacked. The final re-
duced image was built by adding the summed exposures col-
lected during the same night, taking into account of their re-
spective total exposure time. It consists of 1900 × 1900 pixels
corresponding to a total field of view of 6.3′ × 6.3′. The re-
sulting mean seeing, measured from the brightness profile of
individual stars, is found to be around 0.65′′. The central core
of IRC+10216 is measured with a S/N ratio larger than 100
per pixel and the S/N of the envelope at 20′′ from the center is
around 5-6 per pixel. For a more detailed description of this re-
duction procedure see Vandame (2002), and a preliminary pre-
sentation of this image can be found in de Laverny (2003). Due
to the wide field of this image, we have estimated a PSF refer-
ence by using the median average of a set of suitable point-like
stars, that have been, firstly, background-subtracted, centered
with sub-pixel accuracy and normalized. The final FORS1 im-
age (see Sect. 4) has been deconvolved using a Richardson-
Lucy algorithm (Richardson 1972; Lucy 1974). This procedure
has slightly improved the spatial resolution (to less than about
0.6′′) and the S/N ratio (to about 7-8 per pixel at 20′′ from the
center).
3. The CSE innermost regions
The JHKL diffraction-limited images of IRC+10216 are dis-
played in Fig. 1 with a log-scale for the brightness, so that
details of the morphology at all flux levels can be seen. The
labels A to D shown in the H-band image indicate the features
identified and labeled by Haniff & Buscher (1998). The im-
ages have been centered at the central star location estimated
by Murakawa et al. (2005). These authors have performed a
polarimetric study of IRC+10216 in H-band which has inde-
pendently provided a possible central star position, after series
of contradictory hypothesis (see Weigelt et al. 2002 and Tuthill
et al. 2005). Following Murakawa et al. (2005), we have used
clump A as reference to identify the central star position. We
note that their observations were made at the same epoch as the
images presented here.
3.1. Morphologyatdifferentwavelengths
The images exhibit a bright and inhomogeneous structure
which roughly looks like a ring (with a diameter of approxi-
mately 0.5′′) composed of clumps (including clumps A to D)
around an approximately circular depression. The depression
is located at about (0.15, 0.15)′′ from the image center and has
about 6% of the intensity peak. Clumps A and C are the bright-
est features. Clump B appears as an elongated feature, at about
the NE direction from clump C, and clump D seems to be a
more spread out feature. In addition, there is a faint and almost
spherical extended envelope (from ∼ 0.5′′ to more than 1′′ from
the image center), which also appears to have its center in the
ring depression. This depression could thus correspond to an
apparent center of the images. Regarding the central star, it is
located in the fainter SE region of the mentioned ring. Its po-
sition also coincides with a particular elongation in the ring
brightness distribution, well seen in the H-band image. Finally,
we have verified that the faintest structures seen in the J-band
are ghosts, probably due to the light reflected in the NACO op-
tics.
On another hand, a clear difference between the JHKL im-
ages concerns the brightness variations of clumps A to D with
respect to the images peaks. Clump A remains close to the in-
tensity peak in all bands whereas clumps B and D are brighter
at larger wavelengths (from about 10% and 20%, respectively,
of the intensity peak in J to about 100% in L). The bright-
ness of clump C increases more slightly and is always brighter
than 80%. We also note that the brightness difference between
the four clumps strongly decreases with increasing wavelength.
The SE region of the ring, close to the assumed location of the
central star, remains faint, varying from about 8% to 20% of
the intensity peak from J to L. Finally, the extent of the ex-
ternal envelope seems to decrease with increasing wavelength.
Considering its limits at 0.2% of the image peak brightness, we
have calculated its mean extent as being about 4.8′′, 3.5′′, 2.7′′
and 2.0′′ in the J-, H-, K- and L-band, respectively.
The clump brightness variations as well as the extent of
the envelope at different wavelengths may reveal that we
mostly detect, in K & L, the emission of the heated dust,
whereas at shorter wavelengths the scattered stellar emission
becomes more important. This is in agreement with models
of the spectral energy distribution computed for IRC+10216
(see e.g. Mauron et al. 2003). In the K & L bands, the dust
emission is indeed ∼ 100 times larger than the scattered stellar
light which becomes dominant below ∼ 1µm. Therefore, we
can deduce that most of the clumps seen close to the star in
the L-band have approximately the same temperature. On the
contrary, in the J-band optical depth effects could explain the
different brightness of the clumps.
3.2. Temporalvariations
Temporal changes of the IRC+10216 innermost regions have
already been reported by Tuthill et al. (2000), Weigelt et
al. (2002) and references therein. Weigelt et al. (2002) have
estimated approximately linear displacements between clumps
A-C and A-D, and a possible acceleration of 5 mas yr−2 for the
separation A-B. Tuthill et al. (2000, 2005) have identified two
sub-components in clump B: NE1 and NE2 close and far from
clump C, hereafter referred as B1 and B2, respectively. They
have proposed that clumps B1, B2 and D move away from A,
apparently with an uniform acceleration of 3.4 mas yr−2.
4
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
Fig. 1. NACO JHKL images of IRC+10216. Contour levels are 80, 60, 40, 20, 10, 8, 6, 4, 2, 1, 0.5, and 0.2% of the peak surface
brightness. North is up and East is left. The white cross at each image center represents the assumed central star position and
its size is proportional to the error of 0.03" as given by Murakawa et al. 2005. Clumps A to D of the H-band follow the Haniff
& Buscher (1998) clump nomenclatures. The resolution is about 70 mas in the J and H bands, 80 mas in K and 120 mas in L
(represented by the circles at each image corner).
We have also applied on the H-band1 image a Fourier filter-
ing procedure in order to remove the structures of lower spatial
frequencies (see Fig. 2). We have then identified new features,
1 We have selected the H-band image (instead of the K-band most
commonly analyzed) because of its better spatial resolution and con-
siderably better dynamic range. We have verified that the detected
clumps and their estimated positions are similar to those found in the
K-band.
0, B′
1 and B′
in particular, the sub-features B′
2, in clump B. We
found a feature, not identified in previous works, close to the
assumed star position (labeled H). We note that the star posi-
tion assumed in this work could still be discussed and its con-
nection with clump H is very unclear. This clump could be, for
instance, a dust cloud just passing between the star and the ob-
server. Future observations are needed to study the evolution of
this clump with respect to the central star position.
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
5
B2'
B
B0'
B1'
D
H
B
2
1
B
D
C
A
100
50
0
-50
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
0.2
0.0
-0.2
-0.4
-0.6
Relative R.A. (arcsec)
-100
100
50
0
-50
-100
Relative R.A. (arcsec)
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
0.6
0.4
0.2
0.0
-0.2
Fig. 2. H-band map of IRC+10216, where only the highest spa-
tial frequencies of the brightness in the Fourier space have been
kept. The contour levels are 80, 50, 20, 10, 8, 5, 2, 1 and 0.5%
(this minimum level being the estimated noise). The dotted cir-
cle is the assumed star position, as in Fig. 1. Main clumps are
indicated by the labels A to D, and some sub-features by the
smaller labels. The positions of these clumps are shown by the
white crosses. Estimated trajectories (from the results of Tuthill
et al. 2000) for some clumps with respect to clump A are also
shown (see text for details).
Fig. 2 shows estimated apparent trajectories for the previ-
ously detected clumps B1, B2 and D, with respect to A. These
estimations were made by assuming that the clumps move
away from A, as proposed by Tuthill et al. (2000), and by tak-
ing into account of their spatial separations. The solid arrows
represent the displacement of these clumps during the inter-
val time of their observations, i.e. from 1997 January to 1999
April. The dashed arrows show a prediction for the clumps dis-
placements up to 2003 March, by assuming the averaged veloc-
ity of Tuthill et al. (2000). The dotted arrows represent an al-
ternative prediction by assuming the acceleration law proposed
by Tuthill et al. (2000). The error margins are about 10 mas
for the dashed arrows and 30 mas for the dotted ones. We can
see that the previous clumps B1 and B2 are most probably the
current sub-features B′
2, respectively. They are currently
separated by 258 ±20, 394 ±20 and 261 ±20 mas from A. B1
and B2 thus appear to be less accelerated than expected. At
the same time, clump D appears to have moved with a con-
stant velocity. ¿From the clumps separations given by Weigelt
et al. (2002), we have also verified that clump C (currently lo-
cated at 131 ±20 mas from A) appears to be approximately
motionless. Note that choosing clump A as reference could give
the illusion that the clumps escape from it. The clumps motions
are therefore not compatible with the uniform acceleration law
1 and B′
Fig. 3. FORS1 deconvolved V-band image of IRC+10216.
North is up and East is left.
proposed by previous studies, although some accelerations may
exist for clumps B1 and B2. New high angular resolution obser-
vations are needed to disentangle the three-dimensional mor-
phology of the innermost environment of IRC+10216 and to
study the temporal variations of these clumps.
4. External layers of IRC+10216
Fig. 3 shows the deconvolved wide-field V-band image of
IRC+10216. As already shown by MH99-00, we see an ex-
tended halo composed of thin and irregular multiple-shells.
They appear to be non-concentric and azimuthally incomplete.
The CSE is seen due to external illumination by the ambient
Galactic light, scattered by the dust. Since these photons can
easily penetrate into the CSE (their optical depth from the out-
side towards the center being very low), the incomplete shells
do reveal lower densities in some parts of the CSE. The shell
discontinuities can obviously not be caused by some shadow-
ing effects due to more external material.
4.1. Structureoftheexternallayers
To emphasize the shell morphology, we have removed the cen-
tral extended halo by applying the same Fourier filtering pro-
cedure as for the NACO images. We have also removed sev-
eral sources (stars or galaxies) by selecting those with observed
intensities larger than a prefixed threshold. The source pixels
were replaced by averaged values taking into account the local
background and the noise level. The resulting image is shown
in Fig. 4. We have then applied an azimuthal smoothing of 20o
around the center. Although this decreases the spatial resolu-
tion in the azimuthal direction, the resulting map (Fig. 5) shows
6
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
100
50
0
-50
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
-100
100
50
0
-50
-100
Relative R.A. (arcsec)
Fig. 4. Deconvolved V-band image after subtraction of the halo
of the CSE and removing of most stars and galaxies.
100
50
0
-50
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
-100
100
50
0
-50
-100
Relative R.A. (arcsec)
Fig. 5. Schematic map of the shells surrounding IRC+10216
(see text for details).
a clear visualization of the shells, and gives a more realistic and
complete pattern than that presented in previous works.
The shell distribution seen in Figs. 4 & 5 are similar to
those reported before. For instance, the three faint and appar-
ently thick shells at North, located at about 30 -- 40′′, 50 -- 60′′
and 70 -- 80′′ from the center, can also be seen in the CFHT im-
age (MH99-00). However, these shells, being better resolved
in the FORS1 images, appear to be composed of a complex
sub-distribution of thinner ones. Similar thin shells located very
close each other are well seen everywhere in the CSE. For in-
stance, the shells labeled e and f by MH00 (located at dis-
tances to the center of about 55′′ and 58′′, between 347 -- 20o
and 23 -- 53o, respectively) are clearly mergers of complex thin
shell distributions. Another prominent shell located to the S, at
about 15′′ from the center, joins series of slightly less promi-
nent thinner ones distributed toward SE, between 10 -- 30′′ from
center. Moreover, even in the more internal regions, several thin
shells seem to merge in thicker ones between about 4′′ and 20′′.
The whole CSE thus appears to be composed of a complex of
several thin irregular shells that could be identified as thicker
ones in less resolved images. Finally, we note that the separa-
tion between apparently thick shells varies a lot with respect to
the radial direction.
On another hand, a smooth azimuthally radial profile de-
rived by computing the mean of all the pixels found in annuli
0.9′′ thick (see MH99-00 for more details) confirms that the
dust is detected up to about 200′′ (about 5 800 stellar radii).
That corresponds to material ejected about 8 000 years ago
(assuming an escape velocity of 14 km s−1 and a distance of
120 pc), i.e. an important fraction of an interpulse on the AGB.
Actually, we do not see any edge for the dusty envelope and we
are limited by the detector size.
4.2. Thicknessoftheshells
MH00 have proposed that the shells thickness increases with
increasing the distance to the center, in agreement with the nat-
ural expansion of the envelope. To analyze the shell profiles and
to verify their finding, we propose here a new and more accu-
rate method, by taking into account the non-concentric nature
of the shells and the possibility that an apparently thick shell
may be resolved into several thinner ones.
Fig. 6 shows the CSE morphology in a map of narrow2
radial strips. From this diagram, we clearly see the non-
concentric nature of the shells as well as the complexity of their
spatial distribution. For instance, the long and thin shell located
at distances to the center of about 15 -- 20′′, between 220o and
340o from North, has an inclination of about 4o in this diagram
with respect to the vertical axis. The shell located at distances
to the center of about 30 -- 35′′, between 300o and 360o from
North, has an inclination of about 17o. At the same time, there
are shells with opposite orientations, as those found at distances
to the center of about 25 -- 30′′ and 45′′, from 210o to 260o, and
from 200o to 230o, respectively, which make angles of about
−16o and −2o with respect to the vertical axis.
Regarding the profile of the thickest shells, we have care-
fully analyzed the shell labeled d in MH00. Fig. 7(a) shows its
profile by applying the same method as those authors. Fig. 7(b)
shows the profile of the same shell estimated from the more re-
stricted region located at a distance to the center of about 39′′,
between 70 -- 90o, where it appears more regular. We have then
integrated profiles perpendicular to its direction and subtracted
2 Radial sections with thickness of 1 pixel and a rotation step of
0.2o. Interpolations between the original image pixels were applied
for each step.
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
7
Fig. 6. Map of the shells transformed from polar coordinates to a Cartesian representation. Each horizontal section of these maps
represents a narrow radial strip of the FORS1 images, the angles of the strips being with respect to North. Maps to the left, center
and right were derived from Figs. 3, 4 & 5, respectively.
the extended halo contribution. The estimated FWHM of the
profile (a) is about 3.0′′ (as in MH00), whereas it is about 2.6′′
for the profile (b). The error margins are around 0.4′′. Although
both estimates are in agreement within the error bars, a deeper
analysis of this shell reveals that even our profile (b) could be
widened due to a merging of two thinner ones. Indeed, the pro-
files shown in Fig. 7 have two close peaks at offsets of about
±0.5′′, which leads us to suspect that there are two thin shells
close together in this region and, hence, not well spatially re-
solved. We note that this shell was carefully analyzed, this pair
of peaks being identified in every derived profiles. If we decom-
pose the profile (b) in two close shells, their estimated FWHM
are about 1.8′′ ± 0.4′′. In consequence, we have derived several
shell profiles by identifying, as above, well resolved thin shells
at different distances to the center, (see Fig. 8). The profiles (a)
to (d) have good S/N ratios. The profile (e) having a worse S/N
ratio is actually a thin feature composing an apparently thicker
shell which was also detected by MH00. The FWHM of the
two features in the profile (a) and the other four features in
the profiles (b) to (e) are, respectively, about: 1.2′′, 1.6′′, 1.5′′,
1.7′′, 1.6′′ and 1.4′′. The error margins are around 0.4′′. In a
more general way, we have estimated the FWHM of 23 shells
at radial distances from 4′′ to 80′′ (see Fig. 9). We found a
mean FWHM value of 1.6′′, with a standard deviation of 0.3′′.
The minimum FWHM value of 1.2′′ is found for the innermost
analyzed shell, located at about 4′′ from center between 40 --
100o, and also for two shells located at about 9′′ and 11′′ from
the center, between 150 -- 180o and 205 -- 255o, respectively. The
)
%
(
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
(a)
d = 39´´
ang = 55−101o
(b)
d = 39´´
ang = 70−90o
Relative offset (arcsec)
Fig. 7. Comparison of the profile of the shell d estimated as in
MH00 (top panel) with its profile derived by our more complex
method (bottom panel, see text for details). The distances to the
center, d, of the shells and the ranges of their azimuthal angles
from north, ang, are given. The relative intensity is with respect
to the central peak brightness of the original image.
maximum FWHM value of 2.0′′ is found for the shells located
at about 24′′ and 25′′, between 105 -- 135o and 60 -- 85o, respec-
tively. The furthest analyzed shell, located at about 80′′ from
center, between 30 -- 45o, has a FWHM ≃ 1.9′′. The error bars
vary between 0.3′′ and 0.6′′. Thin shells are thus detected in the
whole envelope and even far from center. We therefore cannot
derive a clear increasing relation between the shell thickness
8
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
)
%
(
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
(a)
d = 11´´, 15´´
ang = 150−180o
(b)
d = 18−20´´
ang = 220−330o
(c)
d = 31−35´´
ang = 300−350o
(d)
d = 42−43´´
ang = 200−230o
(e)
d = 55−57´´
ang = 20−30o
Relative offset (arcsec)
Fig. 8. Profiles of some shells considering their non-concentric
nature (see text for details). There are two shells in the panel
(a) and one in the others. The given parameters follow the same
definitions as in Fig. 7. The ranges in the shell distances to the
center, d, are due to their non-concentric nature.
5. Global view of IRC+10216
In order to better understand the possible links existing between
the almost spherical shells and the inner clumps, we describe
here the morphology of the inner CSE from the FORS1 image
together with the NACO data.
Fig. 10 (left panel) shows a closer view of Fig. 4. MH99-00
have detected three structures suspected to be shells in regions
within about 3.1′′ from the center, whereas no such shells are
found in our data, possibly because the HST data have a bet-
ter spatial resolution, despite their lower S/N ratio. The clos-
est identifiable structures are located between ∼3′′ and 16′′
from center. Regarding the core of the nebula, it appears clearly
asymmetric. Two dominant lobes much brighter than the rest of
the envelope lie around the center, making together a direction
of about 22o ±2o with respect to North. The southern lobe being
40% brighter than the northern one. These features likely result
from scattered stellar photons in contrast to the reflected galac-
tic light seen elsewhere in the envelope. Such bipolar morphol-
ogy could be an indication that scattering is more efficient in
the polar direction. It could be roughly reproduced by a simple
model of scattering dust grains in a non-spherical dusty enve-
lope, with evacuated polar regions, around the star, the system
being tilted away from the observer (see e.g. Men'shchikov et
al. 2001). We however note that the main shape of the bipolar
nebula slightly differs from the one reported by MH00 from
their HST image, possibly due to the different spatial resolu-
tion.
Fig. 10 (right panel) gives a representation of
the
IRC+10216 core, by superposing the NACO and FORS im-
ages. The two images were arbitrarily positioned by coincid-
ing their intensity peaks. We are conscious that this assumption
may be crude since the V and IR images result from very differ-
ent physical processes. However, this composite image repre-
sents for the first time both the extended layers and the bipolar
core together with the real size of the inner clumps, and puts
forward the difficulty of finding a link between such small and
large scale morphologies. Firstly, evidence of clumps far from
the center was not found by Huggins & Mauron (2002) in a pre-
vious analysis of the same FORS1 image. Secondly, shells can-
not be identified in the NACO images. We note that, although
the region composed of clumps A to D roughly looks like a
ring, its center (the depression feature) is not compatible with
the star position estimated by Murakawa et al. (2005). Finally,
the bipolar structure detected in V-band is also not clearly iden-
tified in the near-IR.
Fig. 9. Relation between the thickness of the shells and their
distance to the center.
6. Conclusion
and the distance to the center such as that proposed by MH00.
However, the shells found rather close from center could be
resolved into even thinner ones. We therefore could have over-
estimated their thickness. In consequence, either the slope of
the thickness variation with distance proposed by MH00 could
still be valid but with very thin shells close to the star, or the
shell thickness increases much less than that estimated by these
authors.
We have described in this work very high quality images of
the CSE of IRC+10216, from its most inner regions to the
most external ones. In the central arcsec scale of the JHKL
images, sub-arcsec structures (or clumps) identified by other
authors have been recovered about 4 years later. We have also
derived a map of the brightest clumps found close to the core of
IRC+10216. The morphology of these clumps varies strongly
with increasing wavelengths and we propose that the closest
structures have about the same temperature. Furthermore, by
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
9
20
10
0
-10
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
-20
20
10
0
-10
-20
Relative R.A. (arcsec)
)
c
e
s
c
r
a
(
.
c
e
D
e
v
i
t
a
l
e
R
2
1
0
-1
-2
2
1
0
-1
-2
Relative R.A. (arcsec)
Fig. 10. View of the IRC+10216 CSE inner morphology. Left panel shows the closest structures around the center detected in
V-band. Right panel shows the V-band image core, on which has been superposed the NACO H-band image. The overlapping
has been done by assuming that the NACO and FORS peak brightness are found at the same location.
analyzing their apparent relative motion, we cannot confirm the
uniform outflow acceleration previously proposed since only
two bright clumps appear to be accelerated (but at a smaller
rate than that already estimated), whereas others clumps could
have a constant velocity or even no relative motion. At much
larger spatial scales (up to a few arcmin), we present a new map
of the non-spherical incomplete shells characterizing the CSE
of IRC+10216. Owing to the high spatial resolution of our im-
age, most of the thicker shells actually appear to be composed
of thinner elongated ones. Their thicknesses appear rather uni-
formly distributed between about 1′′ and 2′′, regardless of the
distance to the center. Finally, we have combined the NACO
and FORS images in order to provide a more global view of
this CSE and to compare the typical size of the clumps found
very close to the center with the bipolar nebula and with the
much more external shells.
This study has confirmed the very complex nature of the
IRC+10216 envelope with asymmetries already present on the
AGB. Neither the morphology at different spatial scales nor the
motions detected very close to the center can be satisfactory
explained by current models on the mass-loss mechanisms in
AGB stars and their typical time-scales. For instance, Sandin &
Hofner (2004 and references therein) have predicted shell den-
sity distributions not compatible with those observed around
IRC+10216 (see also on the same topic Meijerink et al., 2003).
Another scenario for the formation of the shells in a spheri-
cally symmetric stellar wind has been explored by Soker (2000,
2002). He proposed that these shells could be connected to cool
magnetic spots on the stellar surface. If these spots are more
concentrated near the equator, the mass-loss geometry could
deviate from sphericity and thus favor the formation of shell-
like features and/or clumps.
Moreover, it is interesting to note that the very complex
structures found around IRC+10216 may affect the chemical
composition of its envelope. For instance, the clumps detected
very close to the central core may favor, by their thermody-
namical properties, the formation of the graphite observed in
presolar dust grains (Bernatowicz et al. 2005). Furthermore, the
presence of high density shells in the photochemically active
regions could change the molecular distribution in the envelope
by blocking external UV photons (see e.g. MH00; Brown &
Millar 2003). Then, high contrast shells of complex molecules
may be formed more easily, as confirmed by some millime-
ter observations (see for instance, HCO+, C2H, C4H and HC5N
maps by Gu´elin et al., 2000, and CO maps by Fong et al., 2003).
Finally, future high spatial resolution images of this CSE
are still mandatory in order to better understand the motions
of its clumps (and in particular the clump H superposed on the
assumed central star), their formation/fading, the central star
position and the possible evolution of the external shells as their
three-dimensional morphology.
Acknowledgements. We thank N. Mauron for fruitful and stimu-
lating discussions over all these years and his comments on the
manuscript, and D. O'Brien for proofreading it. The Brazilian agen-
cies CAPES and CNPq are thanked for financial support. P. de Laverny
acknowledges support from the CNRS/INSU (Actions Th´ematiques
Innovantes) and MESR (Jeunes Chercheurs).
References
Appenzeller I., et al. 1998, The Messenger, 94, 1
10
I.C. Leao et al.: The CSE of IRC+10216 from milli-arcsecond to arcmin scales
Bernatowicz, T.J., Akande, O.W., Croat, T.K., et al. 2005, ApJ, 631,
988
Brown, J.M., & Millar, T.J. 2003, MNRAS, 339, 1041
Fong, D., Meixner, M., & Shah, R.Y. 2003, ApJ, 582, L39
Gu´elin M., Lucas R., Neri R., et al. 2000, in IAU Symp 197, p.365,
Minh & van Dishoek eds
Haniff, C.A. & Buscher, D.F. 1998, A&A, 334, L5
Hanuschik R., Amico P. 2000, The Messenger, 99, 6
Huggins, P. J., & Mauron, N. 2002, A&A, 393, 273
de Laverny P. 2003, in Mass-losing pulsating stars and their cir-
cumstellar matter, p. 197, Nakada Y., Honma M., Seki M. eds.,
Astrophysics & Space Sience Library vol. 283, Kluwer Academic
Press.
Lenzen, R., Hartung, M., Brandner, W., et al. 2003, SPIE, 4841, 944
Lucy, L.B. 1974, AJ, 79, 745
Mauron, N., & Huggins, P.J. 1999, A&A, 349, 203
Mauron, N., & Huggins, P.J. 2000, A&A, 359, 707
Mauron, N., de Laverny, P., & Lopez, B. 2003, A&A, 401, 985
Meijerink, R., Mellema, G., & Simis, Y. 2003, A&A, 405, 1075
Men'shchikov, A.B., Balega, Y., Blocker, et al. 2001, A&A, 368, 497
Murakawa, K., Suto, H., Oya, S., et al. 2005, A&A, 436, 601
Richardson, W.H. 1972, JOSA, 62, 55
Rousset, G., Lacombe, F., Puget, P., et al. 2000, SPIE, 4007, 72
Sandin, C, & Hofner, S. 2004, A&A, 413, 789
Soker, N. 2000, ApJ, 540, 436
Soker, N. 2002, ApJ, 570, 369
Tuthill, P.G., Monnier, J.D., Danchi, W.C., et al. 2000, ApJ, 543, 284
Tuthill, P.G., Monnier, J.D., & Danchi, W.C. 2005, ApJ, 624, 352
Vandame, B. 2002, SPIE, 4847, 123
Weigelt, G., Balega, Y., Blocker, T., et al. 1998, A&A, 333, L51
Weigelt, G., Balega, Y.Y., Blocker, T., et al. 2002, A&A, 392, 131
List of Objects
'IRC+10216' on page 1
|
astro-ph/9907282 | 1 | 9907 | 1999-07-21T00:15:08 | The Linear Polarization of Sagittarius A* II. VLA and BIMA Polarimetry at 22, 43 and 86 GHz | [
"astro-ph"
] | We present a search for linear polarization at 22 GHz, 43 GHz and 86 GHz from the nearest super massive black hole candidate, Sagittarius A*. We find upper limits to the linear polarization of 0.2%, 0.4% and 1%, respectively. These results strongly support the conclusion of our centimeter wavelength spectro-polarimetry that Sgr A* is not depolarized by the interstellar medium but is in fact intrinsically depolarized. | astro-ph | astro-ph |
The Linear Polarization of Sagittarius A* II.
VLA and BIMA Polarimetry at 22, 43 and 86 GHz
Geoffrey C. Bower1,2, Melvyn C.H. Wright3, Donald C. Backer3, & Heino Falcke2
ABSTRACT
We present a search for linear polarization at 22 GHz, 43 GHz and 86 GHz
from the nearest super massive black hole candidate, Sagittarius A*. We find
upper limits to the linear polarization of 0.2%, 0.4% and 1%, respectively.
These results strongly support the conclusion of our centimeter wavelength
spectro-polarimetry that Sgr A* is not depolarized by the interstellar medium
but is in fact intrinsically depolarized.
Subject headings: Galaxy: center - galaxies: active - scattering - polarization
1.
Introduction
The compact non-thermal radio source Sgr A* is recognized as one of the most
convincing massive black hole candidates (Maoz 1998). Recent results from stellar proper
motion studies indicate that there is a dark mass of ∼ 2.6 × 106M⊙ enclosed within 0.01
pc (Genzel et al. 1997, Ghez et al. 1998). Very long baseline interferometry studies at
millimeter wavelengths have shown that the intrinsic radio source coincident with the dark
mass has a size that is less than 1 AU and a brightness temperature greater than 109 K
(Rogers et al.
1994, Bower & Backer 1998, Lo et al. 1998, Krichbaum et al. 1998).
Together these points are compelling evidence that Sgr A* is a cyclo-synchrotron emitting
region surrounding a massive black hole. Nevertheless, specific details of the excitation of
high energy electrons, their distribution and the accretion of infalling matter onto Sgr A*
are unknown (e.g., Falcke, Mannheim & Biermann 1993, Melia 1994, Narayan et al. 1998,
Mahadevan 1998).
1National Radio Astronomy Observatory, P.O. Box O, 1003 Lopezville, Socorro, NM 87801
2Max Planck Institut fur Radioastronomie, Auf dem Hugel 69, D 53121 Bonn Germany
3Astronomy Department & Radio Astronomy Laboratory, University of California, Berkeley, CA 94720
– 2 –
We have recently demonstrated that Sgr A* is not linearly polarized at a level of 0.2%
at 4.8 and 8.4 GHz (Bower et al. 1999, hereafter Paper I). This spectro-polarimetric result
excludes rotation measures up to 107 rad m−2. Interstellar depolarization in the scattering
region (Frail et al.
but not completely excluded by these observations. Interstellar depolarization can occur if
the scale of turbulent fluctuations in the scattering medium are on the order of 10−4 pc.
Although this scale is probably too large, it is not fully excluded by observations. The
millimeter polarimetry that we describe in this paper directly addresses the significance of
interstellar depolarization on these scales.
1994, Lazio & Cordes 1998) is unlikely
1994, Yusef-Zadeh et al.
Our recent detection of circular polarization in Sgr A* gives particular relevance to the
question of the level of intrinsic polarization (Bower, Falcke & Backer 1999). Typically,
AGN display integrated circular polarization that is an order of magnitude or more less
than the integrated linear polarization (Weiler & de Pater 1983). This is not only the
consequence of beam dilution. In the case of the VLBI detection of circular polarization
in a compact knot in 3C 279, the circular polarization is less than the co-spatial linear
polarization by a factor of ∼ 10 (Wardle et al. 1998). That is, there are no known regions
in jets with high circular polarization and low linear polarization. Therefore, the presence of
a large circular to linear polarization ratio in Sgr A* is an unsolved and intriguing radiative
transfer problem. We discuss later some of the models that may account for this ratio.
In §2 we present VLA4 and BIMA5 array polarimetry. There is no detected polarization
for Sgr A* at 22, 43 and 86 GHz. In §3 we demonstrate that interstellar depolarization at
these frequencies is extremely unlikely. We consider the consequences of an intrinsically
unpolarized Sgr A* in §4.
2. Observations and Data Reduction
2.1. VLA Observations at 22 GHz and 43 GHz
We observed Sgr A* on 3 February 1997 at 22 GHz and 43 GHz using the VLA The
array was in the BnA configuration. Data were obtained in two 50 MHz wide intermediate
frequency (IF) bands at 22.435 and 22.485 GHz, and 43.315 and 43.365 GHz, respectively.
4The VLA is an instrument of the National Radio Astronomy Observatory. The NRAO is a facility of
the National Science Foundation, operated under cooperative agreement with Associated Universities, Inc.
5The BIMA array is operated by the Berkeley-Illinois-Maryland Association under funding from the
National Science Foundation
– 3 –
The 27-element array was divided into two sub-arrays that observed simultaneously at
22 GHz and 43 GHz. The flux density scale was set by assuming standard flux densities
for 3C 286. Hourly observations of B1730-130 were used to measure antenna-based gain
amplitude fluctuations and to determine the antenna-based polarization leakage terms,
following standard practices. Absolute position angle calibration was not possible due to
errors in the cross-correlation data for 3C 286. All measured position angles were rotated
so that the position angle for B1730-130 was set to 0.
Sgr A* and the compact source B1741-312 were each observed twice an hour for 7
hours. The compact source B1921-293 was observed at 43 GHz once an hour for 4 hours.
Total and polarized intensities in each IF band were measured as the best-fit Gaussian
in the I and P images (Table 1). The quoted errors are rms errors from the fit. We also
report the off-source maximum value in the polarized image, Plim in flux units and plim as
a fraction of the total intensity. A real detection must be more than twice this value to be
believable.
The measured polarizations for Sgr A* are many times the rms image noise, which is
on the order of 0.2 mJy. However, there is a significant contribution from multiplicative
errors. These errors principally derive from variations in the polarization leakage terms
(Holdaway, Carilli & Owen 1992). The effect of the D-term errors is to scatter a fraction
of the total intensity into the polarized intensity map. Typically, at centimeter wavelengths
the VLA can achieve a fractional error of ∼ 0.1% (e.g., Boweret al. 1999a). The smaller
number of antennas and poorer performance of the array at 22 GHz and 43 GHz will lead
to larger fractional polarization errors.
Comparing results between IF bands is not a reliable method for determining fractional
errors. The dominant sources of D-term errors are common to both antennas. Hence, we
see variations between IFs for bright sources that are fully consistent with the thermal
noise.
Two factors indicate that the measured polarization for Sgr A* is an upper limit rather
than a detection. We show in Figure 1 a 43 GHz image of Sgr A* with polarization vectors
overlaid. First, there is large variation in the polarization position angles over the source.
This is also true in the 22 GHz images. Second, the sidelobes and noise peaks are polarized
at a level comparable to the central source. Off-source peaks in the P maps are as large as
the measured polarization. This implies fractional polarization errors of 0.2% and 0.4% at
22 GHz and 43 GHz, respectively.
– 4 –
2.2. BIMA Observations at 86 and 90 GHz
Polarimetric observations of Sgr A* were obtained with the BIMA array (Welch et al.
1996) on three dates, 10 March 1998, 14 March 1998 and 19 December 1998. The array was
in the A configuration producing projected baselines for Sgr A* in the range 20 to 520 kλ.
Continuum bandwidths were 800 MHz in lower and upper IF sidebands centered at 86.582
GHz and 90.028 GHz. Standard antenna amplitude gains were applied.
Each receiver is sensitive to linear polarization. Quarter-wave plates were installed on
all antennas such that the receivers can be switched between linear, right circular (RCP)
or left circular (LCP) polarization. One antenna observed linear polarization continuously,
while the other antennas were switched between RCP and LCP using a Walsh-function
pattern to optimize the visibility coverage in parallel- and cross-hand correlations (Wright
1995, Wright 1996). The data were self-calibrated for both RCP and LCP with respect
to the antenna observing linear polarization. Because RCP and LCP is detected with the
same receiver in each antenna, there is no phase-offset between the parallel hand visibilities.
Hence, the absolute position angle is correctly determined without any further calibration.
For all three observations instrumental leakage was calibrated from observations of
strong unresolved sources. The instrumental leakage is stable to about 0.4% rms. This
implies that the minimum error in the polarization maps will be 0.4%. If variations in the
D-terms are correlated, the error could be over 1% (Holdaway, Carilli & Owen 1992).
For the March observations, we used D-term solutions from spectral-line observations of
the Orion SiO maser on 28 January 1998 and 25 February 1998 (Rao et al. 1998). The
average difference per antenna between the Orion maser D-term solutions is 1.3%, implying
a minimum error in the polarization of ∼ 0.4% if the variations between antennas are
uncorrelated. The average difference between the two Orion maser and calibrator D-term
solutions is similar. This implies that we are not strongly affected by variations in the
D-term solutions over the bandpass. Because solutions were found for a spectral line, they
were available only at a single IF frequency. For the 19 December 1998 observations, we
used solutions found for 3C 273 observed on 21 November 1998 in the C array. These
data showed better agreement between the two IF bands than the solutions found from
interleaved observations of B1730-130. A similar level of variation in the D-term solutions
was found for these observations.
We summarize the total and polarized intensity in Table 2. The reported errors are
estimated from fits to the corrected parallel- and cross-hand visibility data. As is the
case with the VLA data, these are underestimates because they do not take into account
amplitude calibration and polarization leakage term errors. We estimate the total error by
the level of off-source peaks in the polarization maps. These are on the order of 20 mJy,
– 5 –
or 1%, for Sgr A*. This is consistent with the results of Rao et al. , in which the linear
polarization limit is 1.5%. Therefore, we consider the measured polarization for Sgr A* to
be an upper limit of 1%.
In Figure 2 we summarize all upper limits to the polarization of Sgr A* from Paper I
and from this paper.
3.
Interstellar Depolarization
A very large rotation measure (RM) will rotate the position angle of linear polarization
through the observing band. However, bandwidth depolarization is unlikely to occur in
these observations. The maximum rotation measure detectable in the continuum band of
these experiments is 1.3× 106 rad m−2, 8.4× 106 rad m−2 and 4.8× 106 rad m−2 at 22 GHz,
43 GHz and 86 GHz, respectively. The spectro-polarimetric observations in Paper I would
have detected a signal at these RMs if they were present.
We argued in Paper I that the scattering medium will depolarize the source if variations
in the RM lead to a phase change of π radians. The required RM variations at 22 GHz,
43 GHz and 86 GHz are 1.8 × 104 rad m−2, 6.4 × 104 rad m−2 and 2.7 × 105 rad m−2.
The known variations in the RM in the Galactic Center region (Yusef-Zadeh, Wardle &
Parastaran 1997) are not sufficient to depolarize Sgr A* at 4.8 GHz and 8.4 GHz (Paper I).
Therefore, we must only consider whether the depolarization conditions could arise in the
scattering medium around SgrA*.
The angular broadening of images of masers near the Galactic Center and Sgr A* is
most likely associated with the ionized skins of molecular clouds. The ionization mechanism
is either photo-ionization by hot stars (Yusef-Zadeh et al. 1994) or contact with diffuse,
hot gas (Lazio & Cordes 1998). There are two relevant length scales for the structure
of these scattering screens: the thickness of the ionized skins, lskin ∼ 10−4 pc which was
derived by Yusef-Zadeh et al.
(1994); and the outer scale of the turbulent spectrum of
electron density fluctuations within these skins, l0 ∼ 10−7 pc which was derived by Lazio
& Cordes (1998). The small outer scale in relation to the skin depth suggests that these
layers may contain many independent turbulent cells. The small angular scale of these cells,
l0/8 kpc ∼ 0.02 mas, means that they can depolarize a linearly polarized signal owing to
their random Faraday rotations. The rms RM along independent lines of sight through a
single skin will depend on l0qlskin/l0. This rms will be about some mean if the magnetic
field is uniform in the skin or about zero if the field is random. If our line of sight traverses
N skins, then the equivalent path length for the rms RM estimation is L = √Nlskinl0. This
– 6 –
path length is less than 10−5 pc for N < 10 scattering screens.
The constancy of maser image anisotropy over ∼< 10 arcsec angular scales suggests
that the average perpendicular to the line of sight magnetic field imbedded in these skins
is uniform over physical scales of ∼< 1 pc (Yusef-Zadeh et al.
significant fraction of the size of molecular clouds in the Galactic Center region. Hence, the
variations on greater scales may be the result of scattering by physically distinct regions.
This uniformity then requires the rms RM to be about some mean RM (with contributions
from density alone) rather than about zero (with contributions from density and field).
1999). This scale is a
We show now that for L as large as 10−4 pc, depolarization in the scattering medium
and energy equipartition between the magnetic field and particle energy require that either
or both the electron density and magnetic field strength exceed the peak values measured
in the Galactic Center region. These two conditions require
ne = 7.3 × 104 cm−3 RM4
2/3L−2/3
−4 T −1/3
4
and
B = 1.6 mG RM4
1/3L−1/3
−4 T 1/3
4
,
(1)
(2)
where RM4 is the rotation measure in units of 104 rad m−2, L−4 is the length scale in
units of 10−4 pc and T4 is the electron temperature in units of 104 K. Mehringer et al.
(1993) showed that ionized densities in H II regions are significantly less than 105 cm−3 on
arcsecond scales. Magnetic field strengths measured with OH masers in dense molecular
regions are on the order of a few milliGauss (Yusef-Zadeh et al. 1999).
At 22 GHz and assuming T4 = 1, we find B ≈ 2 mG and ne ≈ 105 cm−3, which exceeds
the observed upper limit on electron density. At 86 GHz, B ≈ 5 mG and ne ≈ 7× 105 cm−3.
For the case of L ∼ 10−7 pc, depolarization of the 22 GHz radiation requires B ≈ 15 mG
and ne ∼> 107 cm−3. The case is much worse at 86 GHz. Increasing the electron temperature
does not allow depolarization: it leads to lower electron densities but higher magnetic fields.
Therefore, we consider it extremely unlikely that Sgr A* is depolarized by the interstellar
medium.
These electron densities correspond more closely to what we expect from a sub-parsec
accretion flow onto Sgr A* (Melia 1994, Melia & Coker 1999, Quataert, Narayan & Reid
1999). As Melia and Coker show, densities in excess of 105 cm−3 appear at radii less than
∼ 0.01 pc. We demonstrated in Paper I that this can easily lead to very high RMs and
that depolarization will occur if the accretion region is sufficiently turbulent. However, the
detailed character of the accretion region is not well-known. The geometry, volume filling
factor and degree of turbulence are poorly constrained.
– 7 –
4. An Intrinsically Weakly Polarized Sgr A*
The degree of linear polarization in AGN typically rises with frequency. Aller, Aller &
Hughes (1992) showed that in their flux-limited sample ∼ 40% of AGN have polarization
fractions less than 1% at 4.8 GHz while ∼ 10% of the same sample have polarization
fractions less than 1% at 14.5 GHz. All sources in the sample have detected polarization
fractions greater than 0.2% at 14.5 GHz. This includes 3C 84 which has an average
polarization fraction at 4.8 GHz of 0.03 ± 0.01%. A polarization increase with frequency
can be explained by the high RMs present in some radio cores (Taylor 1998), the increased
prominence of shocked regions and the decreased synchrotron opacity (Stevens, Robson &
Holland 1996). We note that a flux-limited sample of this kind is biased towards powerful,
beamed sources which may have different polarization properties than weaker unbeamed
sources. The polarization properties of these weaker sources are not well-studied due to
their low flux densities. There is no high-frequency polarization study of a volume-limited
sample for weaker sources. However, Rudnick, Jones & Fiedler (1986) did observe a sample
of "weak" cores with flat spectra. They found that even at 15 GHz many of these sources
were unpolarized at a level of ∼ 1%.
The absence of linear polarization in Sgr A* from 4.8 GHz to 86 GHz can be explained
with the presence of thermal electrons or with significant magnetic field cancellation. The
thermal electrons may be outside the emission region (in the accretion flow, as discussed
above, but not in the scattering medium) or may be coincident with the emission region.
This latter case is appealing because it may be able to account for the presence of circular
polarization through the conversion of linear polarization to circular polarization (Bower,
Falcke & Backer 1999, Pacholczyk 1977, Jones & O'Dell 1977).
Magnetic field cancellation could occur as the result of a tangled field or a circularly
symmetric field orientation. The former is typically assumed to depolarize radio jets. This
requires for Sgr A* that the emission region consist of (70/0.2)2 ≈ 105 independent B-field
cells. The latter case may arise if the emission originates in a quasi-spherical inflow (e.g.,
an ADAF model). Magnetic field cancellation is an unlikely depolarization mechanism if
the circular polarization is intrinsic to the source (e.g., Wilson & Weiler 1997). However,
if the circular polarization arises from interstellar propagation effects (Macquart & Melrose
1999), then magnetic field cancellation is a possible explanation. In this case, the absence
of linear polarization argues against a strong shock origin for the total flux variability in
Sgr A* (Wright & Backer 1993, Falcke 1999). Total flux variability in AGN comes about
from the presence of shocks which order the magnetic field and accelerate particles in the
relativistic jet leading to linearly polarized emission (Marscher & Gear 1985).
We have shown here that Sgr A* is not linearly polarized to the current limits of
– 8 –
instrumental sensitivity at 22, 43 and 86 GHz. The possibility is remote that Sgr A* is
externally depolarized. However, the linear and circular polarizations are unique to Sgr
A*. Explaining that relationship may reveal significant details for the emission region and
environment of Sgr A*.
This work was partially supported by NSF Grant AST-9613998 to the University of
California, Berkeley. HF is supported by DFG grant Fa 358/1-1&2.
REFERENCES
Aller, M.F., Aller, H.D. & Hughes, P.A., 1992, ApJ, 399, 16
Bower, G.C. & Backer, D.C., 1998, ApJ, 496, L97
Bower, G.C., Backer, D.C., Zhao, J.-H., Goss, M. & Falcke, H., 1999, ApJ, in press, Paper I
Bower, G.C., Falcke, H. & Backer, D.C., 1999, ApJ, submitted
Falcke, H., Mannheim, K. & Biermann, P. L., 1993, A&A, 278, L1
Falcke, H., 1999, ASP Conf. Series, H. Falcke et al., eds., in press
Frail, D.A., Diamond, P.J., Cordes, J.M. & van Langevelde, H.J., 1994, ApJ, 427, L43
Genzel, R., Eckart, A., Ott, T. & Eisenhauer, F., 1997, MNRAS, 291, 219
Ghez, A., Klein, B.L., Morris, M. & Becklin, E.E., 1998, ApJ, 509, 678
Holdaway, M.A., Carilli, C.L. & Owen, F., 1992, VLA Sci. Memo, 163
Krichbaum, T.P. et al., 1998, A&A, 335, L106
Jones, T.W. & O'Dell, S.L., 1977,ApJ, 214, 522
Lazio, T.J.W. & Cordes, J.M., 1998, ApJ, 505, 715
Lo, K.Y., Shen, Z.-Q., Zhao, J.-H. & Ho, P.T.P., 1998, ApJ, 508, L61
Macquart, J.-P. & Melrose, D., 1999, in preparation
Mahadevan, R., 1998, Nature, 394, 651
Maoz, E., 1998, ApJ, 494, L181
Marscher, A.P. & Gear, W.K., 1985, ApJ, 298, 114
Mehringer, D.M., Palmer, P., Goss, W.M. & Yusef-Zadeh, F., 1993, ApJ, 412, 684
Melia, F., 1994, ApJ, 426, 577
Melia, F. & Coker, R., 1999, ApJ, 511, 750
– 9 –
Narayan, R., Mahadevan, R., Grindlay, J.E., Popham, R.G. & Gammie, C., 1998, ApJ, 492,
554
Pacholczyk, A.G., 1977, Radio Galaxies (Oxford: Pergamon Press)
Quataert, E., Narayan, R. & Reid, M.J., 1999, ApJ, 517, L101
Rao, R., Crutcher, R.M., Plambeck, R.L. & Wright, M.C.H., 1998, ApJ, 502, L75
Rogers, A. E. E. et al., 1994, ApJ, 434, L59
Rudnick, L, Jones, T.W. & Fiedler, R., 1986, AJ, 91, 1011
Stevens, J.A., Robson, E.I. & Holland, W.S., 1996, ApJ, 462, L23
Taylor, G.B., 1998, ApJ, 506, 637
Wardle, J.F.C., Homan, D.C., Ojha, R. & Roberts, D., 1998, Nature, 395, 457
Weiler, K.W. & de Pater, I., 1983, ApJS, 52, 293
Welch, W. J. et al.1996, PASP, 108, 93
Wilson, A.S. & Weiler, K.W., 1997, ApJ, 475, 661
Wright, M.C.H., 1995, BIMA Memoranda #43
Wright, M.C.H., 1996, BIMA Memoranda #48
Wright, M.C.H. & Backer, D.C., 1993, ApJ, 417, 560
Yusef-Zadeh, F., Cotton, W., Wardle, M., Melia, F. & Roberts, D., 1994, ApJ, 434, L63
Yusef-Zadeh, F., Roberts, D.A., Goss, W.M., Frail, D.A. & Green, A.J., 1999, ApJ, 512,
230
Yusef-Zadeh, F., Wardle, M. & Parastaran, P., 1997, ApJ, 475, L119
This preprint was prepared with the AAS LATEX macros v4.0.
– 10 –
C
E
S
C
R
A
0.6
0.4
0.2
0.0
-0.2
-0.4
-0.6
0.6
0.4
0.2
0.0
ARC SEC
-0.2
-0.4
-0.6
Fig. 1.- A total intensity image of Sgr A* with polarization contours overlaid from IF 1 at
43 GHz. The scatter in the polarization vectors over the compact source and the strength
of the off-source polarization indicate that the polarization peak is an upper limit. The
total intensity contours are -1%, 1%, 3%, 10%, 30% and 90% of the peak intensity. A
polarization vector one arcsecond long represents a polarized intensity of 33.3 mJy/beam.
The synthesized beam is shown in the lower left corner.
– 11 –
Fig. 2.- Upper limits to the linear polarization of Sgr A*. Broad band observations are
indicated with an arrow. Spectro-polarimetric observations are indicated with an arrow and
a cross.
– 12 –
Table 1. Polarized and Total Flux from VLA Continuum Observations at 22 GHz and 43
GHz
Source
IF
I
(Jy)
P
(mJy)
Plim
(mJy)
p
(%)
plim
(%)
χ
(deg)
B1730-130
B1741-312
Sgr A*
B1730-130
B1741-312
B1921-293
Sgr A*
1
2
1
2
1
2
1
2
1
2
1
2
1
2
11.342 ± 0.072
11.748 ± 0.064
0.663 ± 0.001
0.668 ± 0.001
1.053 ± 0.001
1.061 ± 0.001
11.512 ± 0.004
11.482 ± 0.004
0.476 ± 0.002
0.479 ± 0.002
14.154 ± 0.032
14.161 ± 0.032
1.074 ± 0.001
1.073 ± 0.001
22 GHz
301.0 ± 2.3
281.4 ± 2.4
29.8 ± 1.0
28.2 ± 0.5
2.0 ± 0.5
1.7 ± 0.5
43 GHz
227.0 ± 3.2
219.5 ± 4.5
28.6 ± 1.2
29.3 ± 1.5
115.6 ± 12.4
89.6 ± 15.3
3.4 ± 0.9
3.7 ± 0.8
21.5
26.7
2.9
3.0
2.1
2.2
9.5
11.3
3.5
3.1
12.2
17.9
2.6
2.5
2.65 ± 0.02
2.40 ± 0.02
4.49 ± 0.15
4.22 ± 0.08
0.20 ± 0.05
0.16 ± 0.05
1.96 ± 0.03
1.91 ± 0.04
6.01 ± 0.25
6.12 ± 0.31
0.82 ± 0.09
0.63 ± 0.11
0.32 ± 0.08
0.34 ± 0.08
0.19
0.23
0.44
0.45
0.20
0.21
0.08
0.10
0.74
0.65
0.09
0.13
0.24
0.23
. . .
. . .
−35.9 ± 1.3
−38.9 ± 0.7
−34.4 ± 10.1
−12.3 ± 11.9
. . .
. . .
9.4 ± 1.7
7.0 ± 2.1
10.2 ± 4.3
8.2 ± 6.9
−29.3 ± 10.7
91.7 ± 8.8
– 13 –
Table 2. Polarized and Total Flux from BIMA Continuum Observations at 86 GHz
Source
IF
I
(Jy)
P
(mJy)
Plim
(mJy)
p
(%)
plim
(%)
χ
(deg)
10 March 1998
3C273
3C454.3
Sgr A*
3C273
3C454.3
B1730-130
Sgr A*
3C273
3C454.3
B1730-130
Sgr A*
1
1
1
1
1
1
1
1
2
1
2
1
2
1
2
25.250 ± 0.340
5.564 ± 0.095
1.715 ± 0.053
23.180 ± 0.368
4.816 ± 0.105
2.824 ± 0.020
1.352 ± 0.065
18.400 ± 0.045
18.250 ± 0.083
5.828 ± 0.119
5.646 ± 0.069
2.841 ± 0.001
2.754 ± 0.002
2.374 ± 0.016
2.422 ± 0.018
742.8 ± 13.2
47.8 ± 6.4
17.3 ± 3.2
39.3
28.3
23.1
14 March 1998
726.3 ± 2.7
39.1 ± 13.8
71.3 ± 1.2
9.5 ± 7.0
19 December 1998
44.0
37.8
67.4
35.6
1166.1 ± 4.7
1136.2 ± 7.0
291.7 ± 10.4
338.7 ± 19.2
77.5 ± 21.1
85.5 ± 7.7
22.8 ± 5.7
22.9 ± 3.6
94.1
75.3
34.3
30.4
33.8
34.0
20.9
19.6
2.94 ± 0.05
0.86 ± 0.11
1.01 ± 0.18
3.13 ± 0.01
0.81 ± 0.29
2.53 ± 0.04
0.70 ± 0.52
6.34 ± 0.03
6.23 ± 0.04
5.00 ± 0.18
6.00 ± 0.34
2.73 ± 0.74
3.11 ± 0.28
0.96 ± 0.24
0.95 ± 0.15
0.15
0.51
1.35
0.19
0.78
2.39
2.63
0.51
0.41
0.59
0.54
1.19
1.23
0.88
0.81
−2.1 ± 0.5
40.9 ± 3.8
−32.6 ± 7.5
−3.5 ± 0.1
50.5 ± 10.1
50.8 ± 0.5
−12.1 ± 29.9
−38.5 ± 0.1
−38.5 ± 0.2
−85.2 ± 1.0
−84.8 ± 1.6
10.8 ± 7.8
14.9 ± 2.6
26.5 ± 10.1
−39.1 ± 6.4
|
astro-ph/0403499 | 1 | 0403 | 2004-03-20T21:26:17 | Far-Ultraviolet Spectroscopy of Star-Forming Regions in Nearby Galaxies: Stellar Populations and Abundance Indicators | [
"astro-ph"
] | We present FUSE spectroscopy and supporting data for star-forming regions in nearby galaxies, to examine their massive-star content and explore the use of abundance and population indicators in this spectral range for high-redshift galaxies. New far-ultraviolet spectra are shown for four bright H II regions in M33 (NGC 588, 592, 595, and 604), the H II region NGC 5461 in M101, and the starburst nucleus of NGC 7714, supplemented by the very-low-metallicity galaxy I Zw 18. In each case, we see strong Milky Way absorption systems from H2, but intrinsic absorption within each galaxy is weak or undetectable, perhaps because of the "UV bias" in which reddened stars which lie behind molecular-rich areas are also heavily reddened. We see striking changes in the stellar-wind lines from these populations with metallicity, suggesting that C II, C III, C IV, N II, N III, and P V lines are potential tracers of stellar metallicity in star-forming galaxies. Three of these relations - involving N IV, C III, and P V - are nearly linear over the range from O/H=0.05--0.8 solar. The major difference in continuum shapes among these systems is that the giant H II complex NGC 604 has a stronger continuum shortward of 950 A than any other object in this sample. Small-number statistics would likely go in the other direction; we favor this as the result of a discrete star-forming event ~3 Myr ago, as suggested by previous studies of its stellar population. (Supported by NASA grant NAG5-8959) | astro-ph | astro-ph | Far-Ultraviolet Spectroscopy of Star-Forming Regions in Nearby
Galaxies: Stellar Populations and Abundance Indicators 1
Department of Physics and Astronomy, University of Alabama, Box 870324, Tuscaloosa,
William C. Keel
AL 35487; [email protected]
Jay B. Holberg
Lunar and Planetary Laboratory, University of Arizona, Tucson, AZ 85721;
[email protected]
Patrick M. Treuthardt
Department of Physics and Astronomy, University of Alabama, Box 870324, Tuscaloosa,
AL 35487
Received
;
accepted
Submitted to the Astronomical Journal
4
0
0
2
r
a
M
0
2
1
v
9
9
4
3
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1Based on observations made with the NASA-CNES-CSA Far Ultraviolet Spectroscopic
Explorer. FUSE is operated for NASA by the Johns Hopkins University under NASA
contract NAS5-32985.
-- 2 --
ABSTRACT
We present FUSE spectroscopy and supporting data for star-forming regions
in nearby galaxies, to examine their massive-star content and explore the use
of abundance and population indicators in this spectral range for high-redshift
galaxies. New far-ultraviolet spectra are shown for four bright H II regions in
M33 (NGC 588, 592, 595, and 604), the H II region NGC 5461 in M101, and
the starburst nucleus of NGC 7714, supplemented by the very-low-metallicity
galaxy I Zw 18. In each case, we see strong Milky Way absorption systems
from H2, but intrinsic absorption within each galaxy is weak or undetectable,
perhaps because of the "UV bias" in which reddened stars which lie behind
molecular-rich areas are also heavily reddened. We see striking changes in
the stellar-wind lines from these populations with metallicity, suggesting that
C II, C III, C IV, N II, N III, and P V lines are potential tracers of stellar
metallicity in star-forming galaxies. Three of these relations - involving N IV, C
III, and P V - are nearly linear over the range from O/H=0.05 -- 0.8 solar. The
major difference in continuum shapes among these systems is that the giant H
II complex NGC 604 has a stronger continuum shortward of 950 A than any
other object in this sample. Small-number statistics would likely go in the other
direction; we favor this as the result of a discrete star-forming event ≈ 3 Myr
ago, as suggested by previous studies of its stellar population.
Subject headings: galaxies:
individual (M33, NGC 7714, M101, I Zw 18) --
galaxies: stellar content -- galaxies: abundances -- ultraviolet: galaxies
-- 3 --
1.
Introduction
The history of galaxies is, in large part, the history of star formation. Massive stars
play key roles both as highly visible tracers of star formation and as players in altering
surrounding star formation and both energy and chemistry of the interstellar medium. It
is these stars which dominate the observed properties of actively star-forming galaxies.
The massive part of stellar populations is most clearly observed in the ultraviolet, where
their energy distributions peak and competing light from cooler stars is minimal, as long as
the foreground extinction allows escape of enough of this radiation. In such environments,
ultraviolet studies of star-forming regions have proven fruitful in understanding these
populations. The recent opening of the far-ultraviolet window, between Lyman α and the
Lyman limit, for deep observations, allows study of massive hot stars in a range where they
fully dominate the spectrum. This relatively narrow band contains an embarrassment of
spectral riches, with numerous lines from stellar winds as well as interstellar material both
atomic and molecular. These include the strong and highly-ionized lines of O VI and the
unique ability to measure cold H2. In addition, this piece of the spectrum is accessible for
high-redshift galaxies, at least for composite samples where the Lyman α forest can be
averaged adequately, allowing the possibility of direct comparisons of stellar populations
over a large span of cosmic time.
The very sensitivity of the far-UV light to star formation and reddening makes it
a purer probe of some properties of star-forming regions than observations at longer
wavelengths. Since only short-lived stars contribute significantly in the far-UV range, the
details of star-forming history should matter only for very brief bursts (such as might be
found in individual H II regions, but are less likely on galaxy scales). This makes the far-UV
spectrum more sensitive to the stellar population itself than to its history. Furthermore,
although the extinction is high, its differential effect across the far-UV band is modest, and
-- 4 --
paradoxically the effective reddening to the stars we see is smaller than found at longer
wavelengths. In observing stars intermingled with highly structured dust distributions, the
"picket-fence" effect (Heisler & Ostriker 1988) means that most of the stars are so reddened
as to make no significant contribution in the deep ultraviolet; all the stars we see are
only lightly reddened. This also reduces the effects of the forest of H2 absorption features
because of the mixing of molecular gas and dust.
Observational data on nearby galaxies, before the Far-Ultraviolet Spectroscopic Explorer
(FUSE), were limited to a handful of star-forming systems. Four starburst galaxies were
observed using HUT on Astro-2 (Leitherer et al. 1995, 2002), largely to measure escaping
radiation in the Lyman continuum, which provided initial data for comparison with models
based on stellar spectra dating back to Copernicus. The strongest features fall into two
blends near 970 (Ly γ + C III) and 1030 A (O VI + Ly β + C II). A Voyager 2 observation
of M33, with some spatial resolution in one direction, was analyzed by Keel (1998), showing
that its far-UV continuum is virtually identical to those of the powerful starbursts, and
that NGC 604 is bluer in this range than the overall disk. These data also suggested
significant Lyman α emission, with spatial profile suggesting an origin in the diffuse ISM
rather than giant H II regions. Ironically, until the availability of FUSE observations, the
richest information on galaxies shortward of Lyman α came from objects at high redshift,
particularly composite spectra of Lyman-break galaxies (Steidel et al. 2001, Shapley et al.
2003). To enable comparisons between local, well-studied star-forming systems and these
powerful, young objects, we have undertaken a series of FUSE observations of star-forming
regions innearby galaxies. We present here the analysis of these spectra in the context of
their stellar populations and systematic changes with metallicity. A companion paper (Keel,
Shapley, & Steidel 2004) considers the comparison with composite spectra of Lyman-break
galaxies.
-- 5 --
2. Observations
2.1. Object selection
The FUSE targets were selected for known high UV flux and hence low reddening,
and to span a range of metallicity. The regions targeted in M33 include the brightest H II
regions in the mid-UV, and span a radial range within the disk from 1.6-4.5 kpc (see Fig. 1).
They also include examples with a single dominant star cluster and with multiple clusters
or more diffuse associations (also illustrated in Fig. 1 using HST data at 1700 A ). None of
the objects in M33 or M101 achieves "super star cluster" (SSC) status, if we follow common
usage in requiring not only high luminosity and stellar mass (Melnick, Moles & Terlevich
1985) but that the stars be concentrated into a single clump on 10-pc scales as noted by
such studies as that of Meurer et al. (1995). While NGC 604 has the right luminosity, its
stars are widely spread throughout a 100-pc region including multiple clumps (Hunter et
al. 1996). In the more distant systems, the FUSE aperture samples multiple regions; Fig.
2 shows the bluest archival HST imagery with the aperture superimposed. In NGC 7714,
there are ≈ 10 luminous SSCs, comparably luminous in the near-UV, within the aperture,
more clearly shown in the inset to Fig. 2. In each of these cases, essentially the entire
star-forming region fits within the aperture, an important desideratum in comparing these
observations with global measurements of distant galaxies.
For each of these objects, there are abundance measurements from the traditional
optical emission lines, which we use as an widely-applicable tracer of the abundances of
recently-formed stars as well. The M33 objects span much of its disk's abundance gradient,
while NGC 5461 and the nucleus of NGC 7714 were selected to sample lower and higher
metallicity. We also analyze the summed archival FUSE spectrum of I Zw 18 from Aloisi et
al. (2003), as a comparison with the lowest known gas-phase abundances.
-- 6 --
Table 1 includes our adopted values from the literature for [O/H], the best-measured
of the abundances due to the optical lines from multiple ionization stages of oxygen and its
importance as a coolant.
2.2. FUSE Spectroscopy
The FUSE optical system and detectors are described by Moos et al. (2000). Four
primary mirrors are used to feed independent detectors optimized for subsections of
the far-UV band; maintaining the coalignment of these optical systems is an important
operational issue. Each detector has a distinct wavelength calibration, but for our purposes,
we are not pushing the resolution limit of FUSE and can combine the various data segments
in 0.1-A pixels.
As shown in Figs. 1 and 2, the FUSE pointing positions were set at the midpoint of
the stellar distributions from available UV images. The large 30" aperture was specified
for the M33 and M101 H II regions at the outset, both for flux integrity in alignment of
the four instrumental channels and to include essentially all their starlight, and adopted for
NGC 7714 as well when the performance penalty for using a smaller aperture became clear
during the mission. The requested center coordinates and total exposure times are listed in
Table 1.
Of these objects, NGC 604 is the brightest in the far-UV range by a factor 10, and will
thus play a continued role in our knowledge of stellar populations in this spectral range. In
each of these spectra, the useful resolution is limited by the signal-to-noise ratio, dictating
the wavelength binning to detect features of interest. For NGC 604, we can work at 0.1-A
binning, for a velocity spacing typically 30 km s−1. Only in this object does it (marginally)
matter that it is not a point source, with starlight coming from most of the aperture area.
-- 7 --
This smearing along the dispersion axis contributes a broadening of about 60 km s−1.
The spectra were processed through the FUSE pipeline; we inspected the two-
dimensional spectra to verify data quality and background subtraction, and to note such
effects as the "worm", a shadowing of the detector by a repeller wire, near the red end of
some spectra. Each of our targets was observed during two orbits, so we could compare
each orbit's summed spectra as a check on the errors at each wavelength. Loss of flux
due to the "worm" does occur in some of the spectra around from 1170 -- 1180 A, a range
which does not affect any of the features we analyze. Similarly, there are spurious emission
features near 1044 and 1169 A which are second-order scattered solar features, which do
not seriously confuse features of interest. For clarity of illustration, we interpolate across
terrestrial airglow features in Ly β, Ly γ and O I at the highest resolution before further
averaging and display.
We also use the summed archival spectrum of I Zw 18 described by Aloisi et al. (2003),
who kindly provided it in electronic form. We compare the spectra of NGC 604 and I Zw
18 at 0.1-A resolution in Fig. 3, showing foreground absorption from H2, atomic absorption
lines in the foreground and in these galaxies, and prominent stellar-wind lines.
2.3. Supporting Data
We draw on additional data from a variety of sources for these well-observed objects.
IUE spectra in the large 10 × 20-arcsecond apertures give mid-UV spectra in roughly
matched apertures; we have collected spectra from the archive, rejecting those with
obviously discrepant pointings or flux levels from the combined spectra.
HST imaging is invaluable in examining the stellar populations, resolving the brightest
stars in each of these associations. Archival WFPC2 data exist for each of the M33 regions
-- 8 --
in F170W plus longer-wavelength filters, so we can assess the stellar statistics at least
into the mid-ultraviolet. The best UV images of NGC 604 were taken in support of a
slitless-spectroscopy program, using the STIS NUV-MAMA detector working at 1800-2700
A .
Each of these stellar collections ionizes a substantial surrounding region. Hα data (as
in Bosch et al. 2002) show structure extending 200-800 pc. All four of the M33 regions
fit their description of an evolved H II region, one in which star formation has been long
enough to generate extensive filaments and loops driven by stellar energy input; they
estimate ages > 4 Myr for all except NGC 604, at 3 Myr, for the current episode of star
formation (matching the conclusions of Gonzalez Delgado & Perez 2000).
Stars which contribute strongly in the far-UV have spectral types B0 and earlier, for
ages up to 15 Myr. However, the continuum slopes of the shorter-lived O stars vary only
slightly across the far-UV range, so the continuum offers little signature of the recent
star-forming history (Robert et al. 2001). To this point, the far-UV line behavior is not
well enough calibrated to infer the star-forming history independently.
3.
Interstellar Absorption Features
The far-UV range is rich in narrow absorption lines from the ISM, both ionic and
molecular. The spectrum of NGC 604 is especially complex, with over 100 detected and
identified transitions. Foreground H2 in the Milky Way is so prominent in all these objects
that it must be carefully accounted for in identifying intrinsic absorption features. For
the strongest Fe II lines, we detect not only the M33 ISM (at about -200 km s−1), but
the foreground high-velocity cloud structure near -370 km s−1 described by Wakker et al.
(2003).
-- 9 --
3.1. Atomic features
Numerous atomic absorption features are seen in these spectra, with stronger
absorption in most cases from foreground Milky Way gas. Their strengths in the integrated
spectra will reflect the distribution of gas toward the stellar associations weighted by
contribution to the far-UV flux; for these extensive distributions of stars, derived column
densities should be regarded as characteristic values.
3.2. Molecular hydrogen
The Lyman and Werner band systems of H2 are at once an advantage and a nuisance in
far-UV spectroscopy, detected in almost every line of sight crossing a significant distance in
our own galaxy's ISM. A total of 56 such lines are individually detected in NGC 604. The
H2 features in these M33 data have also been analyzed by Bluhm et al. (2003), who found
that H2 absorption at the velocities appropriate for disk gas in M33 is weakly detected
in NGC 588, 592, and 595, at column densities N(H2)= 1016 − 3 × 1017 cm−2, but only
an upper limit ≈ 1015 could be derived for NGC 604, despite the higher data quality and
comparable column densities from atomic lines, with all the values approximate due to
the lines being too weak for independent measurement of the Doppler-width b parameter.
Bluhm et al. also present instructive simulations on the difficulty of deriving unique column
densities when the background source consists of stars spanning a range of column density
and position within the spectroscopic aperture. For example, the nondetection in NGC 604
might result from patchy extinction, correlated with the H2 distribution, and the far-UV
background flux being dominated by those stars with the shallowest H2 absorption. This
will act in addition to any physical effects related to the radiation field and shocks in the
vicinity of such an active star-forming region. These regions are certainly associated with
large concentrations of molecular gas; all three of them within the survey area of Engargiola
-- 10 --
et al. (2003) are associated with molecular clouds (NGC 592, 595, and 604 with their clouds
40, 32, and 8, respectively). Their H2 masses are estimated in the range 1.2 − 4 × 105 solar
masses.
In spite of these issues in physical interpretation of molecular absorption, the empirical
results are directly comparable to what we may see in high-redshift systems, as this spectral
region becomes accessible for at least composite study in Lyman-break galaxies (Shapley et
al. 2003). We can use both the sets of H2 line strengths seen from Galactic gas in these
spectra, and synthetic absorption spectra, to ask what signatures of molecular absorption
remain at modest spectral resolution. The least confusion with other spectral features
for modest H2 column densities occurs for blends of features near 1005 and 1072 A. The
wavelengths and shapes of these blends depend on the spin-level populations in the gas,
generally shifting to longer wavelengths for warmer gas (McCandliss 2003).
We have re-examined the NGC 604 spectrum for evidence of associated H2 absorption.
The ability to see such features at low redshift is limited largely by blending with foreground
Galactic features, since the spacing between multiplet members provides ample opportunity
for overlap. For atomic absorption species, we find a difference between foreground and M33
absorption typically 220 km s−1, which gives wavelength shifts very close to the spacing
of low-order members of the H2 multiplets near 963, 982, 1002, and 1014 A . As a result,
only the bluest members of these multiplets could be securely detected from gas near the
velocity of NGC 604 in the presence of the much stronger foreground gas. To guard against
unrecognized atomic features, we also require that a putative NGC 604 feature not be
seen in the I Zw 18 spectrum at the corresponding emitted wavelength. Compared to 56
H2 lines seen from foreground Milky Way gas, only two potential detections resulted for
NGC 604. These are the transitions at rest wavelengths 1001.69 (5, 4 → 13, 1) and 1008.38
(2, 3 → 8, 0) at equivalent widths of roughly 0.02 and 0.08 A , respectively. The 1008 line,
-- 11 --
with better S/N ratio, appears substantially broader, with a profile compatible with that
of the other narrow features. If accurate, these detections are insufficient to derive the
absorbing column density, lacking information on the intrinsic Doppler broadening and
level populations. If we adopt the "optimum" b-value 10 km s−1 for the NGC 588 and 592
detections from Bluhm et al., and incorporating the H2OOLS template models described by
McCandliss 2003, we derive N(H2)=3 × 1017 cm−2 from this single transition and assuming
the same mix of level populations as in the foreground gas. Hotter molecular gas could give
comparable absorption at column densities several times lower, close to the Bluhm et al.
limit.
Lines arising in the states J = 0 − 3 are seen in the foreground Milky Way gas, well
below the saturated level that drove Rachford et al. (2002) to use detailed profile fitting. As
in their work,we also find significant changes in derived N(H2) between various absorption
bands. We follow them in obtaining a characteristic kinetic temperature from the level
populations in J = 0, 1 as
T01 =
74 K
log N(0) − log N(1) + 0.954
.
(1)
For the NGC 604 foreground H2 spectrum, we obtained T01 = 113 K, at the high end of
values seen by Rachford et al. for dense and self-shielded environments but reasonable for
the more diffuse "intercloud" ISM. The populations in J = 0, 1 are taken to represent a
thermal (collisional) temperature, with "excess" absorption from higher J representing
radiative excitation, so that only the lowest levels are useful in deriving the thermal
environment of the molcular gas.
Similar issues affect the strongest H2 lines in I Zw 18, albeit with less overlap from
Milky Way features because of the larger velocity shift. From the stacked spectrum, Aloisi
et al. (2003) set an upper limit of N(H2) = 5 × 10−14 cm−2 against the emerging continuum.
-- 12 --
4. Stellar Content
4.1. Stellar winds
Most of the stellar signature in far-UV spectra occurs in the wind lines, which dominate
this spectral range thanks to the numerous resonance transitions of metals. Both the wind
lines and the evolutionary tracks when winds are important should have strong metallicity
dependences, since the radiation pressure driving the winds acts largely through the opacity
of heavy elements. This behavior has been shown, albeit sometimes in complex ways,
upon comparison of stellar spectra from the Milky Way and Magellanic Clouds. Leitherer
et al. (2001) examined wind lines in the mid-UV, notably C IV, from SMC to Galactic
abundances, confirming a significant metallicity dependence in its strength but noting
than ionization states other than the dominant ones can behave differently as secondary
abundance effects shift the ionization balance in the winds (as seen in Si IV, which is
stronger in LMC stars than the Milky Way). These changes are most pronounced for the
brighter luminosity classes. Among the most luminous stars, those in the Magellanic Clouds
have smaller wind velocities even for comparable depth (e.g. Kudritzki & Puls 2000).
While photospheric lines in the far-UV range are weak enough to scale directly with
abundances, they are either extremely weak or overlain by wind lines. Lamers et al. (1999)
suggest that line blanketing is a more secure route to the stellar abundances. From the
comparison by Robert et al. (2001), Si IV λ1122/8 are high-excitation photospheric lines
that should serve as useful indicators of the stellar population, independent of winds.
In interpreting the O VI profiles from galaxy-scale systems such as NGC 7714,
Gonzalez Delgado et al. (1998) caution from HUT spectra that interstellar absorption from
large-scale outflows can significantly overlap wind lines, specifically between Ly β and O VI
λ1032.
-- 13 --
Even FUSE-based spectral syntheses in the far-ultraviolet extend only to 1000 A
(Robert et al. 2003), so we continue our empirical approach in comparing systems of various
properties to seek differences linked to composition or history. We can, however, be guided
by available syntheses of some of the prominent blends of features, such as the Ly β-O VI -
C II regions from 1025-1038 A (Gonzalez Delgado et al. 1997). They find that Ly β and C
II absorption arise in B stars, at the cool end of far-UV contributors, while O VI P Cygni
profiles come from the most massive stars. High spectral resolution is crucial to separating
the interstellar absorption contributions in each case; for Ly β, hydrogen column densities
NH > 1021 cm−2 lead to blending of stellar and interstellar components. Robert et al.
(2001) show that O VI is very weak at SMC metallicities, and that the youngest population
can be diagnosed from the presence of C II λ1176 and C IV/N IV λ1169 which are specific
to hot O stars. At Magellanic Cloud abundances, the S IV λ1063/73/74 lines have a wind
contribution only from supergiants, and P V λ1118/25 exhibit similar behavior.
To show the some of the spectral differences seen with metallicity, Fig. 4 compares the
spectra of NGC 604 and I Zw 18, now with the strongest Galactic H2 features removed
by fitting Gaussians or Voigt profiles (for the stronger lines), and plotted in the emitted
wavelength frame. Fig. 5 is a similar comparison of NGC 7714 with NGC 604, illustrating
the yet stronger features seen at near-solar metallicity.
The wind lines seen in these systems may be described as follows. For the fainter
objects with short exposures (NGC 588, 592, 595, 5461, 7714) the data have been smoothed
by typically 0.7 A in making these assessments.
NGC 604 (Fig. 4): Combining both members of the O VI doublet to reduce confusion,
there is a broad wind absorption reaching 2800 km/s and probably blending with Ly β.
O VI λ1037, C III λ977, and N II λ 1083 have nearly black cores. A distinct detached
absorption may be present in C III from 1600-2000 km/s. There may be a broad wind
-- 14 --
feature in N III to ≈ 3000 km/s. P V is weak except for possibly photospheric components
near zero velocity. The emission sections of P Cygni profiles are strong in O VI λ1037, N II
λ1082, and N III λ991.
NGC 588/592/595: These spectra have wind features as much like one another as their
S/N ratio can tell, both individually and when averaged. Compared to NGC 604, the cores
of C III, O VI, and N II are shallower, with residual intensities 10-40%. As noted below,
the continuum level in NGC 604 is higher than in any of the other systems below ≈ 950 A.
NGC 5461: a trough occurs in N IV to about 1000 km s−1. There is a single well-defined
trough in C IV to 600 km s−1. No feature is obvious in N III or S III. Broad absorption is
seen in both O VI lines to about 1000 km s−1.
I Zw 18: There is at most weak O VI absorption between 1700-2800 km s−1, seen only
in the λ1037 line. P Cygni emission is absent in O VI, N II, and N III by comparison with
NGC 604. Essentially no blueshifted troughs occur for N III. Two features may be present
for C III from 0-1200 and 1700-2000 km/s, but blending with saturated interstellar C II is
an issue. P V is very weak.
NGC 7714: This nucleus has the strongest wind features in our study. Both O VI
lines are strong, extending to 900 km −1. The P V lines at λλ1118, 1125 are broad and
blueshifted suggesting a wind contribution, which is plausible for non-supergiants at its
metallicity. N III 991 may have a broad wind component blending with the interstellar O I
lines at about 1000 km s−1. The core of C III is blended with O I but shows a wind trough
to beyond 1000 km s−1.
For high-redshift objects, it will be difficult or impossible to separate wind and
interstellar contributions to some of these lines. For purposes of comparison, we have
generated simple equivalent-width values with respect to the adjacent (pseudo)continuum
-- 15 --
for lines in clean parts of the spectrum, typically spanning a 5-A line region, and tabulate
these in Table 2. Airglow emission contaminated the fainter M33 regions (combined in
the table as "M33 avg") too strongly for a reliable measurement of N III λ991. For the N
II and N III lines we list both the absorption and net equivalent widths (in parentheses),
for use when the components are not resolved. We include values for C IV λ1548, 1550
from summed IUE low-dispersion spectra, as these values were taken with a large enough
aperture to sample most of the stellar population in a way much like the FUSE aperture.
Errors on the FUSE value are typically 0.2 A for such broad features, as reflected in the
limits for some lines in I Zw 18, while the C IV errors are closer to 1 A. Each of these
lines shows a strong trend with emission-line metallicity (Fig. 6), showing that they do
in fact have potential use as abundance indicators. The wind lines measure stellar values
directly, in contrast to the large regions of the interstellar medium sampled by emission-line
techniques. These relations are quite sensitive in the sub-solar regime of particular interest
for the evolution of galaxies at z > 3. The equivalent widths measured for N IV, C III, and
P V vary almost linearly with abundances. Our results indicate that these strong features
can be used as metallicity indices for high-redshift stellar populations, in regimes for which
the optical emission lines lie in infrared bands of high atmospheric emission and absorption,
and can be a valuable tool for approaching the chemical evolution of galaxies in the range
z = 3 − 4.
For ease of use, we present least-square quadratic fits to the data in Fig. 6, along
with the derived constants needed to invert these fits with line equivalent widths as the
independent variables. These interpolation curves are shown in Fig. 6, and the numerical
values are found in Table 3. The O/H ratio is in solar units as found from emission-line
analysis (as cited in Table 1, updated to the "new" solar abundance scale with 12 + log
(O/H)=8.69), and all equivalent widths are in A in the emitted frame. For each line,
the data have been approximately fitted in the form EW = a1 + a2(O/H) + a3(O/H)2.
-- 16 --
Similarly, an inverse fitting function for metallicity derived from each line is obtained from
the quadratic formula, with constants tabulated as (O/H) = c1 + c2(c3 + (c4EW))1/2. The
values are based on least-squares fits with equal weights, slightly modified in two cases to
keep the fit monotonic across the metallicity range of our sample. For [N II], we adopt a flat
value EW=1 A for (O/H)<0.5 solar. These forms are for interpolation purposes, and their
limitations appear from the fact that they do not all approach zero line strength at zero
metallicity. Crudely speaking, wind-dominated lines would have a quadratic dependence on
metallicity until saturation sets in, since the mass-loss rate and fraction of the mass in the
right ionic state each depend on metallicity, while mostly photospheric lines (such as P V)
are more nearly linear in strength with the abundances.
Strictly speaking, these relations apply only to galaxies with a long (even if episodic or
weak) history of star formation, since the elements involved in these lines come from very
different stellar processes. For example, oxygen should be enriched rather quickly, coming
from massive stars, while carbon will take longer coming from intermediate masses, with
the dominant sources nitrogen still somewhat ambiguous. At best, these relations could
be taken seriously for single elements in high-redshift galaxies, and can in fact be used to
test for the differential enrichment history of such systems, in much the same way that the
behavior of Si IV and C IV lines with redshifts has been used to infer such differences by
Mehlert et al. (2002), who derive a relation between C IV EW and metallicity consistent
with our IUE analysis included in Fig. 6.
4.2. Far-UV spectra and the stellar mix
Beyond the stellar-wind lines, the greatest difference among all these far-UV spectra is
the flux excess in NGC 604 from about 912-940 A, above what is seen in any of the other
objects either of higher or lower metallicity. Small-number statistics in the massive stars
-- 17 --
might be expected to be more important for the less rich systems, but in M33 it is these
which have the same spectral shape as the more luminous and distant systems, leaving
NGC 604 as the odd one out. Previous work on its stellar content indicates that there was
a strong peak in its star-forming history about 3 Myr ago, recent enough to affect the mix
of stars contributing in this spectral range. We have examined additional UV data on the
stellar content of these M33 H II regions to see how NGC 604 might be different.
The stellar population in NGC 604 has drawn considerable attention, in being one
of the brightest star-forming complexes in the Local Group while offering a dramatic
structural contrast to the dominant, compact stellar cluster in 30 Doradus. Bruhweiler,
Miskey, & Smith Neubig (2003) combined mid-UV WFPC2 images with a wide-slit STIS
spectrum to extract individual spectra of the brightest members. They note that the ten
most UV-luminous stars will dominate the integrated spectral features from NGC 604,
and that two of these are located ≈ 0.3 magnitude above the usual 120-solar-mass limit.
Specifically, the "top ten" stars contribute 46% of the total measured flux just longward of
Lyman α. Mid-UV STIS images (central wavelengths 1820-2700 A) obtained in support of
a slitless-spectroscopy program by J. Mais-Apellaniz show, via the finer pixel sampling, that
there are additional spatially resolved companions to each of the brightest stars identified
by Bruhweiler et al., although none so bright as to bias the measured colors or magnitudes.
NGC 604 also contains significant numbers of WR stars, although with a WR/O star ratio
near 0.1 rather than the 0.3 seen in, for example, NGC 595 (Drissen, Moffat, & Shara 1993).
To compare the populations in these H II regions as they affect the integrated far-UV
spectra, we produced color-magnitude arrays for each based on the archival WFPC2 data
in F170W and F555W (as well as intermediate-wavelength data when available). Lacking
imagery in the far-UV band itself, we use the mid-UV properties as proxies to at least
identify the hottest luminous stars. Reddening corrections do not enter for our immediate
-- 18 --
purposes, since we need to know only how many stars contribute to the UV flux and
extinction does not varying greatly between 1700 and 1100 A . Bruhweiler et al. find, for
stars in NGC 604, that the 1100-A extinction is about 1.5 times that at 1700 A. Likewise,
we are interested here only in stars selected from UV flux, so the diagrams sample only
those stars well detected in the mid-UV data. These observational HR diagrams are shown
in Fig. 7. We include only stars without serious crowding issues, measured within 0.3"
radii. The samples include stars to about m170 = 19 on the STMAG scale (based on flux
per unit wavelength), although the completeness varies considerably even at m170 = 18
because of differences in crowding; for NGC 604, the cumulative counts with flux suggest
statistical completeness only above m170 = 17.0. The brightest stars in the rich population
of NGC 604 are 1.5-2 magnitudes brighter than found in any of the other associations.
The color range among UV-bright stars in NGC 588 is smallest, extending from the blue
envelope near m170 − m555 = −3.5 redward only to -2.0, while all the other regions have
the diagram populated to m170 − m555 = 0, and in NGC 604 to 2.4. This may in part be
a reddening issue, since the stellar distribution in NGC 588 is more compact, with less
scope for differential reddening across the association, than the others. Indeed, dust lanes
are prominent in the continuum images of NGC 604 in both the optical continuum and
UV/optical colors; it is clear observationally that the UV reflects only the least-extinguished
stars in this object. The observed color distribution in NGC 604 is rich in the bluest
stars, but these stars are also represented with similar color in NGC 588 and 595. A more
important difference is that the richer population in NGC 604 includes several stars with
F170W magnitude brighter than seen in any of the other clusters. While richness effects
mean that the brighter star-forming regions will have brighter first-ranked stars, the studies
referenced above suggest that NGC 604 has undergone a distinct burst of star formation
about 3 Myr ago, younger than the other regions, an event which is recent enough to leave
its mark on the massive-star population. We now focus on these brightest stars.
-- 19 --
These data allow us to address how much the brightest stars dominate the mid-UV flux
in each case, incorporating the total F170W flux within the FUSE aperture. Fluctuations
in the bright population could affect the integrated spectra of the lower-luminosity,
sparser regions more strongly simply from statistics, but the detailed recent history of star
formation will enter through aging as well. Given the temperature range of these stars
(3 − 4 × 104 K following Bruhweiler et al.), the hottest stars should be more dominant at
the shorter wavelengths of the FUSE data. It is not clear that all the far-UV flux is from
direct starlight; Hill et al. (1995a,b) and Malamuth et al. (1996) showed evidence that a
significant fraction of mid-UV light from similar systems is scattered. The HST F300W
image of NGC 604 shows reflection nebulosity, and smoothed versions of the mid-UV STIS
images match its morphology, indicating that scattering important at shorter wavelengths
as well, as would be expected for a roughly λ−4 Rayleigh behavior. Since the FUSE
aperture is not much larger than the stellar distribution in NGC 604, the widths of narrow
absorption features are not a sensitive test of whether scattering is important at these
wavelengths. We do see a role for scattering around some of the brightest stars in the HST
STIS image at 2400 A , from analysis of the point-spread functions of stars. This is easier
to interpret than the distinct emission and absorption structures, since reflected continuum
can be confused with emission from the weak [O II] doublet near 2471 A . Some of the stars
match the nominal PSF closely, while two of the brightest ones exhibit excess light from
0.3-1.0" from the core. This excess contains as much of 28% of the mid-UV light in the
brightest case. These data leave open the possibility that scattered light is important at the
shorter far-UV wavelengths.
We therefore bracket the total UV flux between the sum of detected stars and the
large-aperture sum. For larger fractions of scattered light, the brightest individual stars are
more dominant, since the total number of stars producing the observed light is smaller.
Cumulative distributions of F170W magnitude are shown in Fig. 8, along with simple
-- 20 --
geometric estimates of corrections for crowding (which are small, but underestimate
the actual effect where stars are more clumped than random within the H II region).
Accumulating the flux from the bright stars, we find that only about 20% of the total flux
within the FUSE aperture at 1700 A comes directly from the stars, less than half of what
Bruhweiler et al. (2003) find at 1200 A within a narrow slit. This may be a sign that
scattering is important, since much of the scattered light we see at mid-UV wavelengths
in the HST imagery is on larger scales than this. In contrast, as a fraction of the light
from detected stars, the brightest ten (a good approximation to the stars brighter than any
found in the other H II regions) contribute about 40%, more in line with the Bruhweiler et
al. results. This also makes sense for these stars being able to affect the overall spectrum.
However, the high temperature needed for the excess component in NGC 604 means that
we are seeing a difference in history rather than simply small-number statistics in what
stars appear at a given time.
4.3. Stellar populations
In general, metallicity will be manifested in the composite spectra both directly,
through photospheric and wind lines, and indirectly, as the evolutionary tracks of stars
change with abundances. Effects on the initial-mass function are too small to see at
the abundances found in the M33 disk, as shown by Malamuth et al. (1996) for some
of the same H II regions we observed. They suggest that the excitation trends seen in
the associated ionized gas result from the different emergent ionizing fluxes for stars at
various metallicities. However, different evolutionary histories are still implied by the strong
changes in wind properties seen with metallicity.
Comparison of the our spectra suggests that very recent events in the star-formation
history do have observable impact in the far-UV, as exemplified by NGC 604. Its mix
-- 21 --
includes light from a greater fraction of hot stars than the other systems, which fits with
the other properties of this region in suggesting a burst so recent that even in the far-UV,
it does not look like constant star-formation (in this case, ages ≈ 3 Myr). In general, the
timescales may be metallicity-dependent; Robert et al. (2003) show that evolved O I/III
and B I/III stars appear later at lower abundances.
5. Summary
We have used FUSE spectra of star-forming regions in nearby galaxies, whose gas-phase
metallicities range from 0.05-0.8 solar, to explore the utility of far-ultraviolet spectra in
measuring the abundances in star-forming galaxies, as well as to probe the massive-star
populations in these galaxies. The absorption lines from radiatively-driven winds prove to
be very sensitive to metal abundance; all six strong and unblended species (including C
IV from archival IUE data) have a strong, monotonic metallicity dependence. For N IV,
C III, and P V, the relation between straighforward equivalent-width values and oxygen
abundance from emission-line spectra is closely linear, suggesting that these lines will be
useful in tracing the chemical history of galaxies from z = 3 − 4, beyond which the Lyman
α forest makes even composite spectra progressively less informative.
The continuum of NGC 604 departs from the uniform shape of the other objects below
950 A. After considering the effects of small-number statistics among the massive stars in
these objects, we conclude that this difference probably traces to a discrete burst of star
formation ≈ 3 Myr ago in NGC 604. This region had been considered by several previous
studies to have hoisted such a burst, on grounds of both morphology of the gas and fitting
of the H-R diagram.
B.-G. Andersson was helpful in understanding some of the issues in scheduling and
-- 22 --
data analysis from FUSE. We thank Alessandra Aloissi and her collaborators for providing
their summed FUSE spectrum of I Zw 18. We also acknowledge the community service
provided by Steven McCandless in making his H2OOLS compilation of data and routines
available. Dick Tipping patiently explained some of the intricacies of the H2 spectrum on
several occasions. This work was supported by NASA through FUSE GI grant NAG5-8959.
We also made use of the MAST archive system in retrieving data from HST, IUE, and UIT.
We thank the referee, Claus Leitherer, for a detailed, expeditious, and helpful critique.
-- 23 --
REFERENCES
Allende Prieto, C., Lambert, D. L., & Asplund, M. 2001, ApJ, 556, L63
Aloisi, A., Savaglio, S., Heckman, T. M., Hoopes, C. G., Leitherer, C., & Sembach, K. R.
2003, ApJ, 595, 760
Bluhm, H., de Boer, K.S., Marggraf, O., Richter, P, & Wakker, B.P. 2003, A&A, 398, 983
Bosch, G., Terlevich, E., & Terlevich, R. 2002, MNRAS, 329, 481
Bruhweiler, F. C., Miskey, C. L., & Smith Neubig, M. 2003, AJ, 125, 3082
Corbelli, E. & Schneider, S.E. 1997, ApJ, 479, 244
de Vaucouleurs, G. 1959, ApJ, 130, 728
Drissen, L., Moffat, A. F. J., & Shara, M. M. 1993, AJ, 105, 1400
Engargiola, G., Plambeck, R.L., Rosolowsky, E., & Blitz, L. 2003, ApJS, 149, 343
Gonzalez-Delgado, R. M., Perez, E., Diaz, A. I., Garcia-Vargas, M. L., Terlevich, E., &
Vilchez, J. M. 1995, ApJ, 439, 604
Gonzalez Delgado, R. M., Leitherer, C., & Heckman, T. 1997, ApJ, 489, 601
Gonzalez Delgado, R. M., Leitherer, C., Heckman, T., Lowenthal, J. D., Ferguson, H. C., &
Robert, C. 1998, ApJ, 495, 698
Gonz´alez Delgado, R. M. & P´erez, E. 2000, MNRAS, 317, 64
Heisler, J. & Ostriker, J. P. 1988, ApJ, 332, 543
Hill, J.K. et al. 1995a, ApJ, 438, 181
Hill, J.K. et al. 1995b, ApJS, 98, 595
-- 24 --
Hunter, D. A., Baum, W. A., O'Neil, E. J., & Lynds, R. 1996, ApJ, 456, 174
Izotov, Y. I., Chaffee, F. H., Foltz, C. B., Green, R. F., Guseva, N. G., & Thuan, T. X.
1999, ApJ, 527, 757
Keel, W.C. 1998, ApJ, 506, 712
Keel, W.C., Shapley, A.E., & Steidel, C.C. 2004, submitted
Kim, M., Kim, E., Lee, M.G., Sarajedini, A., & Geisler, D. 2002 AJ, 123, 244
Kudritzki, R. & Puls, J. 2000, ARA&A, 38, 613
Kuzio, R. E.; Ciardullo, R.; Feldmeier, J. J.; Jacoby, G. H. 1999 AAS 195, .1103
Lee, M.G., Kim, M., Sarajedini, A., Geisler, D., & Gieren, W. 2002, ApJ 565, 959
Leitherer, C., Ferguson, H. C., Heckman, T. M., & Lowenthal, J. D. 1995, ApJ, 454, L19
Leitherer, C., Leao, J. R. S., Heckman, T. M., Lennon, D. J., Pettini, M., & Robert, C.
2001, ApJ, 550, 724
Leitherer, C., Li, I.-H., Calzetti, D., & Heckman, T. M. 2002, ApJS, 140, 303
Luridiana, V., Esteban, C., Peimbert, M., & Peimbert, A. 2002, Revista Mexicana de
Astronomia y Astrofisica, 38, 97
Malumuth, E. M., Waller, W. H., & Parker, J. W. 1996, AJ, 111, 1128
McCall, M.L., Hill, R., & English, J., 1990, AJ, 100, 193
McCandliss, S.R. 2003, PASP, 115, 651
Mehlert, D., et al. 2002, A&A, 393, 809
Melnick, J., Moles, M., & Terlevich, R. 1985, A&A, 149, L24
-- 25 --
Meurer, G.R., Heckman, T.M., Leitherer, C., Kinney, A., Robert, C., & Garnett, D.R.
1995, AJ, 110, 2665
Moos, H.W. et al. 2000, ApJL, 538, L1
Quillen, A. C. & Yukita, M. 2001, AJ, 121, 2095
Rachford, B.L. et al. 2002, ApJ, 577, 221
Robert, C., Pellerin, A., Aloisi, A., Leitherer, C., Hoopes, C., & Heckman, T. M. 2003,
ApJS, 144, 21
Shapley, A.E., Steidel, C.C., Pettini, M., & Adelberger, K.L. 2003, ApJ, 588, 65
Steidel, C. C., Pettini, M., & Adelberger, K. L. 2001, ApJ, 546, 665
Testor, G. & Lortet, M.-C. 1987, A&A, 178, 25
Vilchez, J.M., Pagel, B.E.J., Diaz, A.I., Terlevich, E. & Edmunds, M. G. 1988, MNRAS,
235, 633
Wakker, B. et al. 2003, ApJS, 146, 1
Waller, W. H., Parker, J. W., & Malumuth, E. M. 1996, ASP Conf. Ser. 99: Cosmic
Abundances, 354
This manuscript was prepared with the AAS LATEX macros v4.0.
-- 26 --
Fig. 1. -- Locations and stellar distributions in the regions observed in M33. The background
image is UIT image FUV-0496 (1500 A ), with ellipses corresponding to galactocentric radii
in the inner disk of M33. Each cutout from HST mid-UV images is 45 arcseconds square
with north at the top, with an outline of the FUSE large science aperture at the recorded
position for each observation. These regions range from single, compact associations (as in
NGC 588) to the extended and multiple collections in NGC 592 and 604. The dispersion
direction for these observations runs ENE-WSW; the limiting spectral resolution is set by
the distribution of the far-UV starlight in this direction, which is a factor only for NGC 604.
The images are displayed with an offset logarithmic intensity scale. The ellipses indicate
distance from the nucleus in the disk plane, taking the geometric parameters for this part of
the disk (within the inner non-warped region) from the optical fits by de Vaucouleurs (1959)
and the H I fits of Corbelli & Schneider (1997). We adopt a distance of 850 kpc, in the
middle of the range of distances from Cepheids (Lee et al. 2002), the tip of the red-giant
branch (Kim et al. 2002), and planetary nebulae (Kuzio et al. 1999).
Fig.
2. -- FUSE aperture location and size for the NGC 5461 and 7714 observations,
superimposed over the shortest-wavelength HST observation available. The ACS image of
NGC 5461, from program 9490 led by K. Kuntz does not include the entire FUSE aperture.
An inset shows the multiple luminous clusters in the starburst nucleus of NGC 7714, observed
by Windhorst et al. under program 9124. As for the cutouts in Fig. 1, each image section
spans 45 arcseconds with north at the top.The dispersion direction for NGC 5461 runs SSE-
NNW, while for NGC 7714 it is ENE-WSW.
Fig. 3. -- FUSE spectra of NGC 604 (heavy line) and I Zw 18, shown with 0.1-A pixels in
the heliocentric velocity frame, to reduce clutter from the H2 absorption features. Fluxes
have been scaled up by 1014 erg cm−2 s−1 A−1 for NGC 604, and by 2 × 1014 for I Zw 18. The
numerous H2 features (of which the most prominent are marked by ticks below the spectra)
-- 27 --
are from Milky Way gas. Features in NGC 604 are slightly blueshifted (240 kms−1, about
0.8 Aat 1000 A), while the redshift of I Zw 18 is cz = 751 km s−1 for a typical wavelength
shift of +2.4 A. Atomic interstellar absorption lines are marked at each redshift by vertical
lines above the spectrum, where the shorter line is for I Zw 18. Angled symbols near the
top indicate the zero-redshift locations of important stellar-wind lines. The continuum of
NGC 604 is higher below about 940 A. While the interstellar lines are nearly as strong in I
Zw 18 as the much higher-metallicity disk of M33, the stellar wind features are substantially
weaker. The continuum level in NGC 604 is significantly higher shortward of about 955 A,
and P Cygni emission redward of the λ1037 line is prominent in NGC 604 but not in I Zw
18.
Fig. 4. -- Comparison of the FUSE spectra of NGC 604 and I Zw 18, as in Fig. 3, now with
Galactic molecular absorption removed and both spectra plotted in the emitted wavelength
frame. Wind and interstellar lines are marked as before; some unpatched foreground
absorption remains.
Interstellar absorption features intrinsic to the surrounding galaxies
stand out by matching in both spectra. This comparison shows the difference in both wind
absorption and P Cygni between the abundances of I Zw 18 (O/H about 0.02 solar) and
NGC 604 (0.4 solar). Both effects are clear for O VI, C III, and N II.
Fig. 5. -- Comparison of the H2-corrected spectrum of NGC 604 to NGC 7714, in the emitted
frame as in Fig. 4. Stronger absorption is prominent in the blue wings of O VI and C III, and
in the overall profiles of N II and Si III/IV. The NGC 7714 data have been boxcar-smoothed
by 0.7 A and scaled by a factor 1014.
Fig. 6. -- Equivalent widths of stellar-wind lines in the spectra of star-forming regions. Each
shows a strong metalliity dependence, here quantified using the traditional emission-line
results for O/H. Typical errors for the far-UV lines are ±0.2 A , with the IUE spectra used
for C IV accurate to about ±1 A. The three fainter M33 regions are averaged into single
-- 28 --
points for each transition. The quadratic interpolation functions with coefficients listed
in Table 3 are overplotted as guides, where we take a flat value EW=1 A for N II below
(O/H)=0.5.
Fig. 7. -- Observed color-magnitude arrays for H II regions in M33, derived from archival
WFPC2 images. The two brightest objects in NGC 604 have less certain colors due to
saturation in the F555W images. These are shown in the STMAG system, in which zero
color index corresponds to constant Fλ. Since we are interested in which stars contribute
to the far-UV flux, no reddening corrections have been applied. The stars of interest are so
blue that red-leak corrections in the F170W filter are negligible for our purposes.
Fig. 8. -- Cumulative star counts as observed at 1700 A for the M33 H II regions, including
stars within the FUSE apertures. In each region, the upper curve includes a simple corection
for crowding, made by assuming that the fainter stars are uniformly distributed through the
populated region in each association.
-- 29 --
Object
RA
α2000
Dec
δ2000
cz
Exposure
O/H
source
km s−1
seconds
(solar units)
NGC 588
01 32 45.50 +30 38 55
-174
NGC 592
01 33 12.27 +30 38 49
-162
NGC 595
01 33 33.60 +30 41 32
-178
NGC 604
01 34 32.50 +30 47 04
-226
NGC 5461
14 03 41.30 +54 19 05
298
NGC 7714
23 36 14.0 +02 09 19
2798
5220
3965
7123
7151
5189
6023
I Zw 18
09 34 02.30 +55 14 25
751
95097
0.41
0.48
0.56
0.66
0.68
0.81
0.05
Vilchez et al. 1988
Interpolated
Vilchez et al. 1988
Vilchez et al. 1988
Luridiana et al. 2002
Gonzalez-Delgado et al. 1995
Izotov et al. 1999
Table 1: FUSE Targets and Properties
Note. -- O/H is in solar units, converted when necessary assuming a solar value of 12+log
O/H=8.60 following Allende Prieto, Lambert, & Asplund (2001)
-- 30 --
Object
N IV 955 C III 977
N III 991
N II 1083 P V 1122 C IV 1549
I Zw 18
0.27:
M33 avg
NGC 604
NGC 5461
NGC 7714
1.26
2.17
2.11
3.26
0.47
1.19
1.83
1.56
2.56
< 0.2 (< 0.2)
1.11 (1.02)
... (...)
0.86 (0.24)
1.77 (0.59)
1.50 (1.33)
2.40 (1.87)
2.19 (1.92)
3.68 (2.36)
2.52 (2.52)
0.27
0.66
0.89
0.90
1.01
2.0:
8.4 (7.8)
9.49 (7.97)
7.82 (6.03)
10.1 (10.2)
Table 2: Equivalent Widths of Stellar Wind Lines
Note. -- All values are in A in the emitted frame. Parenthesized values include the
emission component of a P Cygni profile.
-- 31 --
Transition
a1
a2
a3
c1
c2
c3
c4
N IV
N III
N II
C IV
C III
P V
0.30
-0.656
5.2201
0.063
0.0958
-5.8314
20.8804
0.05
-1.079
6.5539
0.082
0.0763
-0.0794
26.2157
1.28
-4.029
7.0447
0.286
0.0710
-19.7753
28.1790
1.14
19.210
-10.8494
0.885
-0.0461
418.6269
-43.3975
0.50
-0.502
3.6120
0.070
0.1384
-6.9750
14.4481
0.22
0.910
0.1037
-4.386
4.8224
0.7356
0.41473
Table 3: Quadratic Fits for Line Strength versus Metallicity
Note. -- Entries are coefficients of forward and inverse quadratic fits as listed in the text,
when equivalent widths are in A and O/H is in solar units.
This figure "fig1.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig2.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig3.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig4.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig5.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig6.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig7.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
This figure "fig8.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0403499v1
|
0708.3099 | 1 | 0708 | 2007-08-22T21:48:03 | A multicolor near-infrared study of the dwarf nova IP Peg | [
"astro-ph"
] | We report the analysis of $JHK_{s}$ light curves of the eclipsing dwarf nova IP Peg in quiescence. The light curves are dominated by the ellipsoidal variation of the mass-donor star, with additional contributions from the accretion disc and anisotropic emission from the bright spot. A secondary eclipse is visible in the $J$ and $H$ light curves, with 2% and 3% of the flux disappearing at minimum light, respectively. We modeled the observed ellipsoidal variation of the secondary star (including possible illumination effects on its inner face) to find a mass ratio of $q = 0.42$ and an inclination of $i = 84.5^{o} $, consistent in the three bands within the uncertainties. Illumination effects are negligible. The secondary is responsible for 83%, 84% and 88% of the flux in $J$, $H$ and $K_{s}$, respectively. We fitted a black body spectrum to the $JHK_{s}$ fluxes of the secondary star to find a temperature of $T_{bb} = 3100\pm500 K$ and a distance of $d=115\pm30$ pc to the system. We subtracted the contribution of the secondary star and applied 3-D eclipse mapping techniques to the resulting light curves to map the surface brightness of a disc with half-opening angle $\alpha$ and a circular rim at the radius of the bright spot. The eclipse maps show enhanced emission along the stream trajectory ahead of the bright spot position, providing evidence of gas stream overflow. The inferred radial brightness-temperature distribution in the disc is flat for $R < 0.3R_{L1}$ with temperatures $\simeq 3500K$ and colors consistent with those of cool opaque radiators. | astro-ph | astro-ph |
A i
ea ifaed dy f he dwaf va eg
T. Ribei
R. Baia
E. T. a afi
V.S. Dhi
R.G.. R e
1
1 ,2
3 ⋆
4
5
1
2
3
4
5
Deaae de Fíi
a Uiveidade Fedea de Saa Caaia Ca Tidade 88040 900 F iaó i SC
Bazi
e ai : iaga. f
.b
Sa Te e
e C ia E i / Cai a 603 a Seea Chi e
i e f Sa
e A i
ai ad Ree Seig aia bevay f Ahe Bx 20048 Ahe 118 10
Gee
e
De. f hyi
ad Ay Uiveiy f She(cid:30)e d She(cid:30)e d S3 7R U
aa
ew G f Te e
e Aaad de Ce 321 E 38700 Saa C z de a a a Sai
Re
eived ???; a
eed ???
ABSTRACT
We e he aa yi f JHKs igh
ve f he e
iig dwaf va eg i ie
e
e. The igh
ve ae diaed by he e iida vaiai f he a d a wih addiia
ib i f
he a
ei di
ad aii
eii f he bigh . A e
day e
ie i viib e i he J ad
H igh
ve wih 2% ad 3% f he (cid:29) x diaeaig a ii igh ee
ive y. We de ed he
beved e iida vaiai f he e
day a i
dig ib e i iai e(cid:27)e
i ie fa
e
(cid:28)d a a ai f q = 0.42 ad a i
iai f i = 84.5o
eaiie. iai e(cid:27)e
ae eg igib e. The e
day i eib e f 83% 84% ad 88% f he
(cid:29) x i J H ad Ks ee
ive y. We (cid:28)ed a b a
k bdy e
he JHKs (cid:29) xe f he e
day a
(cid:28)d a eea e f Tbb = 3100 ± 500 K ad a dia
e f d = 115 ± 30
he ye. We ba
ed
a he fa
e bighe f a di
wih ha f eig ag e α ad a
i
a i a he adi f he bigh
he
ib i f he e
day a ad a ied 3 D e
ie aig e
hi e he e ig igh
ve
ie i he hee bad wihi he
. The e
ie a hw eha
ed eii a g he ea a je
y ahead f he bigh ii
vidig evide
e f ga ea ve(cid:29)w. The ifeed adia bighe eea e diib i i he di
i (cid:29)a f R < 0.3RL1 wih eea e ≃ 3500K ad
ie wih he f
a e adia.
ey wd. a: vae
aa
yi
vaiab e (cid:21) a: idivid a eg (cid:21) a: e
iig (cid:21) ifaed: a
1. d
i
The hae f he e
day a i
aa
yi
vaiab e
CV ad w a x ay biaie XB i de(cid:28)ed by
i R
he e ieia fa
e. de ig he e iida
eg i a e aive y bigh V ≃ 14.5 ag g eid
Porb = 3.8 h dwaf va hwig e
e b ev
ey few h i whi
h he ye i
eae i bighe
by ≃ 2 ag i he viib e. The biay i ee a edge
i
iai i > 80o
a wig f e
ie f bh he whie
vaiai d
ed i he igh
ve f hee biaie by
dwaf/a
ei di
ad he e
day a a we a f
he
hagig ae
f he died e
day a wih bi
he bigh heeafe BS. i
a ad avi e UV
ay hae vide
ai he bia aaee
igh
ve f eg ae diaed by aii
ei
e.g. Wae 1995 ad yie d a eiae f he
ib
i f he ie BS akig i had
ai
i f he e
day a he a igh.
he ye aaee e.g. Wd Cawfd 1986 ad
U ike viib e ad avi e igh
ve whi
h ae
heefe di(cid:30)
a y e
ie aig e
hi e
diaed by eii f he di
ad bigh he
ea ifaed R igh
ve f CV ad XB have
iveigae i a
ei di
e. Szkdy ae
1986 f d evide
e f e iida vaiai i JHKs
igi(cid:28)
a
ib i f he
e
day a. The
igh
ve f eg. Fig e a . 1999 [heeafe F99℄
deeiai f he e
day a (cid:29) x i a e f wave
f d a
ed d b e h d ai afe b
egh a w a die
eiae f i e
a ye ad
a e he ibi iy f ifeig he dia
e
he biay by ea f hei
aa ax. F hee
kwig he
ib i f he e
day a he igh
ve a w e i ae he igh f he a
ei di
a
ig he e iida vaiai f he e
day a f
hei H bad igh
ve. Thei e
ie aig aa yi
idi
ae ha he a
ei di
i
wih a (cid:29)a adia
eea e diib i Tbb ≃ 3000 K i he H bad.
ad a y e
ie aig e
hi e e.g. Baia
Thi ae e a i
dy f eg i he
Seie 1993 iveigae he di
e i he i
JHKs bad. Se
i 2 ee he bevai ad daa
faed.
ed
i. Se
i 3 de
ibe he
ed e ed (cid:28) he
⋆
e iida vaiai ad he a i
ai f 3 D e
ie a
i eia
ig e
hi e he igh
ve afe we ba
he
ib i f he e
day a. The e ae di
ed
Tab e 1. a f he bevai.
i Se
i 4 ad aized i Se
i 5.
2. bevai ad daa ed
i
eg wa beved wih W RCA ghe e a . 1996
a he 4.2 Wi ia e
he Te e
e i a a a
1996
be 26 29 whi e he ye wa i ie
e
e.
The bevai
ie 3 bia
y
e i he H bad
ad a bi e ha 1 bia
y
e ea
h i he J ad Ks
bad. A wee efed i gd weahe wih bigh
b
d. The eeig aged f 1.0(cid:17) 1.6(cid:17).
The ex e ie wee f 1 3 H ad f 3 J ad Ks .
The bevai ae aized i Tab e 1.
We ed a diheig
ed e eiae he
i
b i f he ky ba
kg d ddig he e e
e i a
(cid:28)ve ii ae ae
ee f
e. The
ky eve wa baied f he edia f he (cid:28)ve iage
f ea
h e ad wa he ba
ed f ea
h iage f he
e. A eie f dak iage wih he ae ex e ie
a he
ie
e fae wee baied d ig he a i
eva f ∼ 1 − 2 h. Thee iage wee ed eve
he dak
e f he
hi. Ce
i f (cid:29)a (cid:28)e d e(cid:27)e
wa a efed.
Daa ed
i wa efed ig A T/RAF
1
ae e hey ie. The fae f ea
h diheig
e e
e wee a iged wih a iea
ive
ed e baed
he RAF egie ak ad
bied i
eae he
iga ie ai S/ f he ea ed a. The
Dae
bad
a
. f
E
ie
hae
1
1996
UT
fae
y
e
age
H
H
Ks
J
10/26
10/27
10/27
10/28
1
19:00
1362
30138
0.860.50
30139
0.500.20
20:00
760
30144
0.280.56
23:16
590
30145
0.420.24
19:15
1212
30150
0.150.50
30151
0.500.34
wih ee
he eheei f e.1
J
H
Ks
18
16
14
12
10
18
16
14
)
y
J
m
(
x
u
l
f
12
10
18
16
14
12
10
ieaiy f he dee
wa
e
ed ig he i i
-0.4 -0.2
0
0.2 0.4 0.6 0.8
1
1.2 1.4
ie i he CT a
kage. F xe wee he exa
ed
f he vaiab e ad f a
ai a 7.4′′
7.5′′
Ea f he vaiab e.
h ad
Phase
Di(cid:27)eeia igh
ve age a (cid:29) x divided by
ai added he ifeed
a di
ib i
Fig. 1. igh
ve f eg wih de ed e iida vai
ai a (cid:29) x wee
ed. The igh
ve wee (cid:29) x
a ibaed ig he 2ASS JHKs ze i
a
Sk kie e a . 2006 ad he ab e agi de f he
ai a J = 9.82 H = 9.56 ad Ks = 9.48.
e
e
i wee a ied. The daa wee
hae f ded a
dig a di(cid:28)ed vei f he iea
eheei f W f e a . 1993
Tmid(HJD) = 2445615.4156(4) + 0.15820616(4)· E,
wee Tmid i he ie f ifei
j
i f he e
1
day a. The e ig igh
ve
a be ee i Fig.
1. The H bad daa wee
bied ive he S/N f
he igh
ve. The J ad Ks bad daa d
ve he
f bi f he biay.
3. Daa aa yi
3.1. The whie dwaf e
ie widh ∆φ.
The i
a dy f Wd Cawfd 1986 ead a
whie dwaf e
ie widh i he age ∆φ = 0.086− 0.092.
The
eaiy i eiaig ∆φ aie f he fa
ha
he whie dwaf e
ie ige i eg i vei ed by he
(cid:28) he daa. Vei
a dahed ie ak he ige/ege
hae f he whie dwaf f a e
ie widh f ∆φ =
0.0918
y
e.
age a i de (cid:29)i
keig ee i i
a igh
ve i
BS ige.
de e(cid:28)e he va e f ∆φ we e yed a ie
aive
ed e a ig biay aaee deived f
he (cid:28) f he e iida vaiai Se
. 3.2 ad a yig
he
edig hae (cid:27)e φ0 eeded ake he b
eved whie dwaf id ege fea e
i
ide wih hae
∆φ/2. The
ed e ea
he f he ai f ∆φ φ0 va
e ha yie d he be χ2
ieaive
e
vege a i wih ∆φ = 0.0918
ad φ0 = +0.01
ie wih he e ii f Wd
Cawfd 1986. We e he J bad igh
ve i hi
f he de igh
ve. Thi
e be
a e i hw
ea y he whie dwaf ege
ike he H bad igh
ve. The Ks bad igh
ve wa
ed f hi
ed e be
a e f i we S/ ad
h e
ed ige f he BS e
ie ad by he
ed
ed hae
veage. Be
a e he whie dwaf ege
fea e i ee i bh bad e
ie we ay afe y
1
ex
de he ibi iy ha he beved fea e i d e a
RAF i diib ed by he aia i
a Ay
bevaie whi
h ae eaed by he A
iai f
Uiveiie f Reea
h i Ay
. de
eaive
(cid:29)i
ke/(cid:29)ae f he di
. The igh
ve hw i Fig. 1
wee
e
ed by he deived va e f φ0 i de ake
ageee wih he aia S
ie
e F dai.
id e
ie
i
ide wih hae ze.
3.2. Fiig he e iida vaiai.
Tab e 2. de ed eg aaee
The e iida vaiai i de ed wih he aid f a igh
ve yhei ga whi
h
e he (cid:29) x eied
by a R
he be (cid:28) ig a a i ae
hage wih biay
hae. The fa
e f he R
he be (cid:28) ig a i divided
i a age be f i e. The (cid:29) x eied by ea
h i e i
di(cid:28)ed a
f gaviy ad ib dakeig e(cid:27)e
.
We ed he gaviy dakeig
e(cid:30)
ie f Saa 1989
β = 0.05 ad aded he iea ae ib
dakeig aw f Diaz Cdvé Giéez 1992
I = I0(1 − a(1 − cosγ) − b(1 − √cosγ)),
2
whee γ i he ag e bewee he ie f igh ad he a
he fa
e f he i e ad I0 i he ieiy eied
a he i e fa
e. We a ed aJ = −0.465 aH =
−0.454 aK = −0.448 ad bJ = 1.199 bH = 1.173 bK =
1.066 C ae 1998.
de a
f bh ib ad gaviy dakeig
we eed ad va e f he eea e Tpole ad gav
iy gpole a he e f he a. Uig a
ed e ii a
ha de
ibed by F99 we eed a age f hyi
a y
a ib e va e vayig Tpole f 2800 3200K ad gg
f 3.5 5.0. The e wee ahe ieiive he
hi
e f aaee ex
e f he Ks bad i whi
h he
ga y
vege f va e f Tpole = 3000− 3200K
ad gg = 4.5. We he de
ided ad Tpole = 3000K
g g = 4.5 ad a ea i
iy i
e hee va e ae i
i
q
Fs y
Fs/FT
Is/I0
ps
J
H
84o ± 5o
0.43 ± 0.10
9.9 ± 2.0
(82 ± 5)%
0.03 ± 0.02
(3 ± 2)o
85o ± 3o
0.42 ± 0.20
11.5 ± 2.5
(84 ± 5)%
0.04 ± 0.03
(2 ± 1)o
Ks
87o ± 2o
0.44 ± 0.03
10.1 ± 1.6
(88 ± 4)%
≤ 0.01
(cid:21)
Fs i he (cid:29) x f he e
day ad FT i he a (cid:29) x
ad 88% f he a bighe i he H ad Ks bad e
e
ive y i ageee wih he e f Szkdy ae
1986 ad F99. i efai e a . 2001 (cid:28)ed a 4V ye
a he Ks bad e
f eg (cid:28)d ha he e
day a
ib e 62% f he a igh a ha wave
egh. weve a ied by ai e a . 2005a
b he deh f he abi ie f he e
day a
i CV ae ed
ed wih ee
i aed a f ii a
e
a ye eha by e ahei
e(cid:27)e
. hi
ae ae a
h he deh f he abi ie
wi yeai
a y deeiae he
ib i f he
e
day a he a igh. Thi e(cid:27)e
ay a
f he we
ib i ifeed by i efai e a . 2001.
A eaive y he di(cid:27)ee ifeed
ib i
d be
bee ageee wih he i he iea e Szkdy
a
e e
e f
hage i di
bighe wih ie wih
ae1986; egge 1992; F99.
a igh y bighe di
eadig a we e
day a
The aaee (cid:28)ed by he e iida vaiai
de
ae: a ai q biay i
iai i he eak ieiy
f a a he ie heihee f he e
day a
e aive he ea ieiy a he a fa
e Is/I0
he ieai ag e f he wih ee
he ie
jiig bh a ps a hae ideede addiive (cid:29) x
eve ad he (cid:29) x f he e
day a a hae ze Fs .
The ga iiize he χ2
bewee he beved igh
e aive
ib i a he e
h f he bevai f
i efai e a . 2001. i had e hi ibi iy
be
a e eg i a he ii f dee
i f aae
ae whi e i ie
e
e hwig a yi
a
ae
f 0.2 ag i i hii
a igh
ve. We aa yzed he
AAVS hii
a igh
ve f eg ad f d di
eib e age ha 0.2 ag di(cid:27)ee
e i bighe ae
bewee he e
h f he bevai f F99 i efai e
ve ad he e iida vaiai de f a give e f
a 2001 ad hi wk.
aaee wih a aeba iiizai
hee e e
The (cid:28)ed Ga ia he ie fa
e f he e
a . 1986. A eaive y i i ib e (cid:28) a hae (cid:27)e
day a i gh y
eeed a he 1 i ad give a
ha iiize he χ2
f ea
h e f aaee ee Se
.
eg igib e
ib i he a (cid:29) x i a bad wihi
3.1.
he
eaiie idi
aig ha iadiai e(cid:27)e
ae
The R igh
ve f eg hw
ib i f
igi(cid:28)
a i he R
i .
he
e aa f he e iida vaiai f he e
A i
iai ve a ai diaga f eg
day a. A a (cid:28) e we eved y he hae
i hw i Fig. 2 whee he deived age f va e f
veig he iay ad e
day e
ie f he igh
q ad i f he hee bad ae dei
ed. Thee i gd
ve befe aeig (cid:28) he e iida vaiai. Thi
ageee bewee he e f he hee bad wih
ead had
vege ea ii
i whi
h
he diei f he a ai va e beig
h a e
deeiae he h a hae +0.75 ad veeiae he
h a hae +0.25 i he J ad H bad. The di(cid:27)ee
e
i bighe i age i he J bad (cid:21) whee he ayey
ha he fa e f he deeiai i he J ad
H bad. The be (cid:28) we χ2
va e i baied f he
bad daa. Be
a e hi igh
ve i he aveage f daa
i he e
ie hae
a ed by he BS i age (cid:21) ad dia
f hee bi i ha he
ee hae
ve
ea i he Ks bad whi
h hw a evide
e f
he BS i he e
ie hae. Thi idi
ae ha aii
age highe S/ ad we i(cid:29) e
e f (cid:29)i
keig. The
J ad Ks igh
ve i
de daa f y e bi
eii f he BS
ib e igi(cid:28)
a y he h
wih i
ee hae
veage a h gh he a
k f hae
eeed a hae +0.75. We heefe de
ided i he
hae age [+0.6; +0.9] a we a he iay ad e
day e
ie f he (cid:28)ig
ed e. Wih he ew
veage f he bad daa de a(cid:27)e
he (cid:28) f i
e iida vaiai. ie f hei a e fa e
he e f he Ks bad ae e e iab e i
e y a
ei
i i wa ib e (cid:28) a de igh
ve he
a a f he igh
ve wa ed f he (cid:28). Takig
daa f a bad.
i a
he
ai deived f he ifeed widh
Tab e 2 i he e f he (cid:28)ig
ed e. The
f he whie dwaf e
ie Se
. 3.1 be i i
de igh
ve ae hw a id ie i Fig. 1. e
idi
ae ha he e
day a i eib e f 84%
i = 84.5o
ad q = 0.42. Thi e f aaee i idi
aed
by a (cid:28) ed
i
e i Fig. 2.
80
)
o
82
(
n
o
i
t
a
n
i
l
c
n
I
84
86
H
88
Ks
90
0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7
Mass ratio (q = M2/M1)
Fig. 2.
iai a ai diaga. Dahed ie i
di
ae he e ai f ∆φ = 0.0863 e
ve ad
∆φ = 0.0918 we
ve ad ba
ke he age f i
J
JHKs .
fa
e bighe diib i f he eg a
ei di
i
e
ie a i a 3D fa
e
iig f a 51 x
51 ixe di
gid wih a ha f eig ag e α he ag e
bewee he id ae ad he di
fa
e
veig he
iay R
he be R = RL1 a
i
a i
f 101 ixe a R = Rbs he ifeed adi f he BS.
Wih hi geey hee i eed exa
he bia
h f he igh
ve F99 i
e i
a be a
ed
f by eii f he
i
a i. We a e ha
hi geey a w he di
bighe diib i ex
ed beyd he adia ii f he BS. The adi f
he BS wa deived a f w. We ea ed he BS id
ige/ege hae φbi = −0.027, φbe = +0.983 i he
J bad igh
ve afe
e
ig f φ0 Se
. 3.1. F
he aded biay geey i he ai f φbi, φbe
va e a i a x y ii i he bia ae whi
h
ie y fa a g he ba ii
ea a je
y. The
BS adi i ake a he adi f he
i
e ha ae
h gh hi ii Rbs = 0.58RL1 . The aded biay
geey ad he BS adi ae dei
ed i Fig. 4. F he
e
ie aig de ig y he daa i he hae age
va e f Wd Cawfd 1986. Vei
a ded ie
[ 0.15;0.15℄ wa aa yzed.
ak he age f va e deeied by Wa e a .
2003. Thee bxe idi
ae he 1 − σ ii he de
ived aaee f he JHKs bad daa. A
a
e
i
e a e diad ad a (cid:28) ed
i
e dei
he
We efed e Ca i ai eiae he
eaiie f he e
i R e e a . 1992.
F a give igh
ve a e f 100 ai(cid:28)
ia igh
ve
i geeaed i whi
h he daa i ae ideede y
fa i f Wd Cawfd 1986 Beeka e
ad ad y vaied a
dig a Ga ia diib i
a 2000 ah 1988 ad hi wk ee
ive y.
wih adad deviai e a he
eaiy a ha
i. The igh
ve ae (cid:28)ed wih he e
ie aig
de d
e a e f adized e
ie a. Thee
ae
bied d
e a aveage a ad a a f
Fig e 2 a
ae e wih he i he
he eid a wih ee
he aveage whi
h yie d he
iea e. The biay aaee deived f he e i
ida vaiai i he J ad H bad daa ae i ea
ab y gd ageee wih he f Wd Cawfd
1986 Beeka e a . 2000 ad Wa e a . 2003.
The i f ah 1988 e ie ifeig he whie
dwaf adia ve
iy f he wig f he eii ie
f he di
. Thi e
hi e i e age
eaiie
ai
a y whe hee i igi(cid:28)
a hae (cid:27)e bewee
he e
i
ifei
j
i f he e
day a
aii
a
eaiy a ea
h ixe . The
eaiie b
aied wih hi
ed e ae ed daw he
a f Fig. 3 ad eiae he
eaiie i he de
ived adia bighe eea e diib i Fig. 5
ad he (cid:29) x ai diaga Fig. 6.
The di
ha f eig ag e α i a fee aaee i he
b e. The ey f he e
ie a i a ef
i ga gig he
e
va e f α. veeiaig de
eiaig he di
ha f eig ag e id
e i
ad he beved id e
ie ie a i i i he
ae i
e i he di
ide
e fahe f he 1
eg.
i. Be
a e he ey i a ea ee f he h
Se a ahee de e.g.
z 1979 ad avai
ab e e a a ae e.g. Se
ke e a . 1979 ae iab e
f (cid:28)ig R (cid:29) xe f vey
a eihe be
a e f a
e
a e i ad a ig i he R be
a e
he age f eea e f he gid/a a de i
de
a
e ha ≃ 3300K . Theefe we (cid:28)ed b a
k bd
ie he exa
ed JHKs (cid:29) xe f he e
day a f
eg i de eiae i eea e ad dia
e. The
be (cid:28) i yie d Tbb = 3100 ± 500K ad a dia
e
f dbb = (115 ± 30)pc a ig R2 = 0.4R⊙ Beeka e
a . 2000. The e ae
ie wih he f Szkdy
ae 1986.
e f he e
ie a hee
e ae (cid:29)agged wih
highe ey. Theefe e ay eiae he va e f α
by efig a e f e
i f a a ib e age
f α ad e e
ig he e wih we ey. Si ai
Bge Baia Caa a 2007 hw ha i i ideed
ib e e he ey a a
iei ife he va e
f α b ha he a f we ey deeiae
he
e
α by 1.5− 2.0o
ey (i, q). We
(cid:28)ed hi (cid:28)dig wih
aef i a
deedig he biay ge
i de wih he e
i(cid:28)
geey f eg f whi
h
we (cid:28)d he (cid:27)e be 1.5o
f α we ed he Ks bad igh
ve whi
h ha he
. de ife he va e
yei
e
ie hae ad he a e
ib i f
3.3. 3 D e
ie aig
Afe ba
ig he e iida vaiai
edig
he aded biay aaee f ea
h igh
ve
wih i
eaig e f ∆α = 0.5o
he bia h . We eed a age f α va e bewee
0o− 8o
α = 0.5o
α = 2o
e ey ad highe degee f yey i baied f
. A
dig he abve i ai we aded
f he e
ie aig e
i. Thi i
. The a f w
e
ie aig e
hi e Baia Seie 1993 wee
ie wih
ai f vei
a di
e by
a ied he eid a
ve deive a f he
eye eye feie 1982 Sak 1992 ad é
2.5
2
1.5
1
0.5
0
)
2.5
y
J
2
m
(
1.5
x
1
u
l
0.5
F
0
2.5
2
1.5
1
0.5
0
J
H
Ks
-0.1 0 0.1
Phase
-7
-7
-7
-6.5
-6.5
-6.5
-6
-6
-6
Total
Asymmetric
-100 0 100
)
r
t
s
/
z
H
/
s
/
2
m
c
/
g
r
e
6
-
0
1
x
(
y
t
i
s
n
e
t
n
I
10
8
6
4
2
0
10
8
6
4
2
0
10
8
6
4
2
0
Azimuth (degree)
Fig. 3. ef: The igh
ve afe
e
i f e iida vaiai d ad he be (cid:28) 3 D e
ie aig de
id
ve. Vei
a ded ie ak he ige/ege hae f he whie dwaf. idd e ef: Cedig di
bighe diib i a gaihi
gay
a e. Ded ie hw he R
he be. A id
ie i ve ed
ea
h e
ie a idi
ae he 3 σ
(cid:28)de
e eve he ieiie. The gaihi
ieiy
a e i idi
aed by
he hiza gay
a e ba i erg/cm2/s/Hz/str. idd e igh: Ayei
di
bighe diib i. Ded
ie hw he R
he be he ba ii
ea a je
y ad a di
f adi 0.58RL1. Righ: Di
i bighe
diib i. The x axi idi
ae he azi h wih ee
he ie jiig bh a. A vei
a ded ie ak
he azi ha ii f he bigh . The ded ie hw he 1 σ ii he ieiie.
f a
ide
e f ga ea ve(cid:29)w i eg. The axi
Ga ia 2000 whi
h edi
α ≃ 1.5o − 2.0o
(cid:10)
= 10−11 − 10−10M⊙yr−1
a
ei ae
.
f he eha
ed eii i wave egh deede
ig fahe dwea i J ha i Ks ib y be
a e
igh
ve
edig e
ie a ad di
i
bighe diib i ae hw i Fig. 3. F a bee
f he geive y age a f gaviaia eegy
vi a izai f
e i he di
bighe diib
e eaed i
ii a he ea aa
he he whie
i he ayei
di
e ae a hw. A
dwaf. Thi ai f eha
ed eii a g he ba ii
yei
e i baied by i
ig he di
i a
ea ahead f he BS ii i eii
e f ha ee
e f adia bi ad (cid:28)ig a h ie f
i he
i he dwaf va WZ Sge S i R e 1998; Skide
ea f he we ha f f he ieiie i ea
h bi. The
e a . 2000. S i R e 1998 i ha
h a
ie (cid:28)ed ieiy i ea
h a a e
i i ake a he
ai i be exe
ed a a
e e
e f he ia
yei
di
eii
e. Thi
ed e e
hyddyai
f he ea.
eve he bae ie f he adia diib i whi e evig
A BS i exe
ed f a he iee
i f he ba
a azi ha
e. The ayei
di
e i
he baied by ba
ig he yei
di
f he
igia e
ie a Sai Baia 2006. The ay
ei
di
e a
f 30% 20% ad 16% f
he a (cid:29) x i J H ad Ks ee
ive y.
ii
ea wih he di
e edge. The ba ii
ea
hi he 0.58RL1 di
i a a azi h θ = 15.6o
wih
ee
he ie jiig bh a
2
. The JHKs di
i
diib i Fig. 3 hw a bigh a a azi h f
2
The di(cid:27)ee
e bewee he va e f θ f he aded =0.42
The a hw eha
ed eii a g he ba ii
geey ad f he a eaive =0.5
ae Wd Cawfd
ea a je
y
e he whie dwaf vidig ev
1986 i eg igib e.
q = 0.42 Rd = 0.58RL1
1
0.5
1
L
R
/
y
0
-0.5
-1
-1
-0.5
0
x/RL1
0.5
1
ae di
f a a
ei ae f
= 10−10, 10−11
(cid:10)
ad
10−12M⊙yr−1
a ig M1 = 1.05M⊙ Beeka e a .
2000 ad R1 = 0.0124R⊙ deived ig he whie dwaf
a adi e ai f a ebeg 1972.
The ie di
R < 0.3RL1 hw a (cid:29)a bighe
eea e diib i eii
e f he ee i i
e
e dwaf vae e.g. Wd e a . 1986 1989 wih i
feed eea e f ≃ 3500 a R = 0.1RL1 f he
JHKs bad wihi he
eaiie. The
ie
y f
he ifeed bighe eea e i he hee bad
gge ha he ie di
i i
a y hi
k wih ei
i
e b a
kbdy. The eea e de
eae i he
e di
egi R > 0.3RL1 i eaab e ageee
wih he T ∝ R−3/4
aw i he J ad Ks bad ad a a
we ha w gadie i he H bad. Thee i agia ev
3
ide
e
f highe bighe eea e i he e di
egi i he H bad wih ee
he J ad Ks bad.
f ea
h H bad ex
e w d idi
ae ha he ga i
he e di
egi i a e wih a vei
a eea e
gadie
h ha he H −
fee fee ad b d fee a
iy ii a 1.6 i
ead a e aive i
eae i
ig (cid:29) x i he H bad wih ee
he J ad
Ks bad. Thi i had e
i e wih he e f
Fig. 4. The geey f eg. A (cid:28) ed
i
e ak he
i efai e a 2001 whi
h gge ha he e egi
exe
ed ii f he BS; a e
i
e dei
he b
f he a
ei di
i eg ae i
a y hi e he
eved azi h f axi eii a g he di
i.
e f he beved i e
ie i a i
a y hi
A aw idi
ae he die
i f axi eii f
he bia h i ed fwad 54o
f he azi ha
hhee abve he i
a y hi
k a
ei di
.
A (cid:29) x ai diaga f he a
ei di
f eg i
ii f he bigh . The age
i
e
ed
hw i Fig. 6. The
f he yei
e f
he e
ie a ae ed gehe wih e aihi f
b a
kbdy BB ai e e
e a S ad i
a y
hi eii . The BB ad e
a wee
ed wih he yh/RAF a
kage. The BB ad
(cid:29) xe ae exa
ed by
v vig he ee
ive e
a di
f adi 0.58RL1.
θ ≃ 20o
∆θ ≃ 30o
wih azi ha exe f widh ha f axi
i eaab e ageee wih he edi
ed
wih he ee f
i f ea
h ifaed abad ad
ii f he BS. The ieiy f he de
eae wih
he (cid:29) x ai ae
ed. The S
wee exa
ed
i
eaig wave egh i a
da
e wih he
ed
f Bee Be 1988 ad afed he 2ASS
ig ed
i i he heigh f he bia h . Thee i
hei
ye ig he e ai f Caee 2001.
a a ede
y f he
eid f he ve wad
The
eaiie i he di
ae ie age ad i
he azi h wih i
eaig wave egh.
eae wad he e ad faie di
egi. The
i ieeig
ae he azi ha ii f
eaiie i he
ae diaed by he we S/ f
he BS i he di
i a wih ha f he bia h
axi . The azi h f h axi i θ ≃ 74o
φmax ≃ −0.2 Fig. 1 ≃ 54o
fwad f he azi h
f he BS. Fig e 4 hw a
heai
diaga f eg
he Ks bad igh
ve ad e
ie a. The
f he
yei
ie di
R<0.3 RL1 ae
ie wih he
f
a e adia wih Tcolor ≃ 3000 − 5000 K a
he 1 σ
(cid:28)de
e eve . The di
be
e edde
e
whee he (cid:27)e bewee he azi h f he BS ad ha f
wih i
eaig adi ad a deviae f he BB e
he axi h eii i
ea. A ii a e(cid:27)e
ha
aihi f R > 0.3RL1 . The beved ed i e di
bee ee i he dwaf vae. f BS eii i d
ed
i
ie wih ha ifeed f he bighe
i a h
k bewee di
ad ea ga he axi ei
eea e diib i ad agai gge a H bad
i wi be a a die
i bewee he di
ad
ex
e (cid:29) x f he e di
egi. The di
ve
ea (cid:29)w Wae 1995 .81. F exa e i Y Ca
wad he e ef f he diaga i he die
i
he h axi i di a
ed ≃ 19o
fwad wih ee
ie ha exe
ed f i
a y hi ga. Whi e he
he azi h f he BS Wd e a 1989. The (cid:27)e i
e ba ae age ad he diaga h d be viewed wih
eg i age. ee he azi h f bia h axi
a i he e gge ha he e di
f eg i
i
ide wih he a he ga ea a he BS
a e wih a vei
a eea e gadie.
ii ggeig ha he ia
h
k ay a g he
ea a je
y wih eg igib e i(cid:29) e
e f di
a
eia . We e ha he hae f axi f he bia
3.4. The e
day e
ie
h i he R igh
ve de
i
ide wih he
e i he i
a φ = 0.9 ee Fig. 1 f Wd Cawfd
1986 a a eady i
ed by F99.
Fig e 5 hw he JHKs di
adia bighe e
ea e diib i f a a ed dia
e f d = 115
. Ded ie hw he T ∝ R−3/4
aw f a e eady
Fig. 1 ad 7 hw he de
ve f he e iida
vaiai gehe wih he daa. i
ea f he J ad
H bad daa ha e
ig edi i
veig a
f he igh f he e
day a bewee hae 0.45
3
The beved di(cid:27)ee
e i eea e i a he 1σ ii.
10000
J
J
1000
10000
)
K
(
e
r
u
t
a
r
e
p
m
e
t
-
s
s
e
n
t
h
g
i
r
B
1000
10000
1000
0.01
H
H
Ks
Ks
0.1
R/RL1
1
-150 -50 50 150
azimuth(deg)
egi w d be eeded
he ae a f igh.
Theefe exe
ie vide a gh we ii he
ize f he a e
ig di
. avig hi i id he
be (cid:28) i ae rd(J) = 0.19±0.06 RL1 α(J) = (0.7±
0.4)o
edig a
ai f 2.3± 0.4 e
e
f he e
day a igh ad rd(H) = 0.22 ± 0.03 RL1
α(H) = (1.5± 0.7)o
ai f 3.0± 0.6 e
e f he
e
day a igh ee
ive y f he J ad H bad.
The e
day e
ie ad he be (cid:28) de
ve ae
hw i Fig. 7. The de vide a gd de
ii f
he daa f hae φ < 0.55 b deviae afe he e
ie
be
a e f he a
ed
ib i f he bia
h he igh
ve. A ig a di
adi f 0.58
RL1 hi gge ha a ea he ie ≃ 1/3 f he
We (cid:28)ed a b a
kbdy de he exa
ed J ad H
a
ei di
i a e i eg i ie
e
e.
bad de(cid:28)
i (cid:29) xe a id e
day e
ie (cid:28)d a e
ea e f he
ed ie heihee f he e
day
a f Toc = 2300± 600 K . Afe
e
ig f he gaviy
dakeig e(cid:27)e
hi be
e T ′
oc = 2500 ± 700 K
i
e a he 1σ ii wih he b a
kbdy eea e f he
e heihee f he e
day a deived i e
i
3.2. The ageee bewee he ifeed eea e f
he ie ad e heihee f he e
day a i i
ie wih he
i deived f he e iida vaia
Fig. 5. ef: Di
eea e bighe diib i
i (cid:28) ae y ha i iai e(cid:27)e
he e
day
a g
a e f he yei
di
e. The id
ve
e
ig he (cid:28) ed ae i he adia (cid:28) e
f he di
ad he dahed
ve idi
ae he 1 σ i
i baied wih e Ca i ai. Ded ie
a f eg i ie
e
e ae eg igib e i he R
i . The
ig aea
ed ≥ 13 e
e
f he je
ed fa
e f he e
day a a hae 0.5.
hw a e eady ae di
de f a a
ei
4. Di
i
ae f
= 10−10, 10−11
(cid:10)
ad 10−12M⊙yr−1
a ig
M1 = 1.05 M⊙ ad R1 = 0.0124R⊙. Righ: Di
i
F99 f d a d b e h ed d ai i hei H bad
bighe eea e diib i a g
a e. A ded
igh
ve afe ba
i f he e iida vaiai f
ie idi
ae he azi ha ii f he bigh .
he e
day a ee hei Fig.5. Thee i evide
e
f d b e h ed d ai i daa. Si
e he ha
ig f hei daa wa e
ed by a
ea idei(cid:28)
ai
f he whie dwaf ege fea e hei aa yi ay be af
ad 0.55. E
ie f he e
day a by he di
wee
fe
ed by e i he a ed biay hae. F99 added a
ee i he H bad daa f F99 (cid:21) wh ied ha
hae (cid:27)e f ∆φ = +0.027 hei igh
ve
e
he ha w deh f he e
ie e ie ha a f he
f deviai f he eheei. weve e
ie ea e
ieveig a
ei di
be i
a y hi (cid:21) ad a
e f whie dwaf e
ie ege ie f
eay
i he e
i
daa f Beeka e a . 2000 (cid:21) wh
bevai f eg i ie
e
e idi
ae a hae (cid:27)
ded
ed ha he a
ei di
f eg be a e
e f ∆φ = +0.008 f he e
h f he F99 bevai
a
f he deh f he e
day e
ie. i efai e
Baia e a . 2005. Th he igh
ve f F99 ae y
a . 2001 f d evide
e f a e
ie f he e
day a
by i
a y hi a f he a
ei di
f hei K
eai
a y hifed by δφ = +0.019 wad iive hae.
deed hee i a wad kik i he (cid:29) x a hae ≃ +0.06
bad ie e ved e
y f eg i ie
e
e.
i hei daa be ee i hei Fig.11 whi
h eeig y
A ig
a hea e i ibi hey f d ha he
ed he whie dwaf ege ggeig a hae (cid:27)e
ig i
a y hi ga i ie h 10000 < T <
≃ +0.015 wih ee
he exe
ed whie dwaf ege
20000 ad a
f a ea he e 20 e
hae. Si ai hw ha
h a a hae (cid:27)e be
e f he di
. ee we de he e
day e
ie
wee he daa ad he e iida vaiai de w d
ife he
f he
ed e
day a fa
e ad
be e gh id
e a i d b e h ed d
eiae he adia exe f he a e
ig di
.
ai i he igh
ve afe eva f he
ib i
A ig he biay geey deived i e
i 3.2
f he e
day a. The hae (cid:27)e a di a
e he
we i aed he e
ie f he e
day a by a a e
wh e di
bighe diib i wad he ai ig ide
y ide f adi rd ad ha f eig ag e α ad ea
hed
f he va e f rd ad α whi
h iiize he χ2
f he
f he di
he e
aiig he ga ea ad w d
a
f
h f he (cid:27)e bewee he exe
ed i
(cid:28) he daa. dig hi we ae akig he i ii
i f he BS ad he ii deived f hei e
ie
a i ha he di
i f y a e a
eai
a.
adi ad be
e f y aae heeafe. f a
iy
The de ed e iida vaiai a idi
ae ha he
hage wih adi ad heigh abve he di
id ae
iadiai f he e
day a fa
e i eg igib e i he
e exe
ha a izeab e fa
i f he di
wi aia y
ifaed. Thi e i i aae
adi
i wih he
ai he igh f he e
day a ad a age
f Davey Sih 1992 ad Wa e a 2003. Deie
2
1.5
1
BB
2000K
)
s
K
(
F
/
)
H
(
F
0.5
K0
MS
4000K
5000K
0.1RL1
0.4RL1
M0
M6
3000K
9500K
4000K
2500K
HI
15
14
13
12
11
)
y
J
m
(
x
u
l
F
J
0
0
0.5
1
1.5
2
F(J) / F(Ks)
15
Fig. 6. F x ai diaga. A id ie idi
ae he
f b a
k bdy eie. aive eea e a g he
e e
e ae idi
aed by abe . A dahed ie hw he
f i
a y hi eiig ga wih hee eea
e abe ed. A ded ie idi
ae e a ai e e
e
wih e e
a ye abe ed. Fi ed ae
e
ed by id ie idi
ae he
f he yei
di
e. The e ba eee he adad deviai
wih ee
he ea va e i ea
h adia bi. abe
ak he adi i i f RL1 .
)
y
J
m
(
x
u
l
F
14
13
12
11
H
0.3
0.35
0.4
0.45
0.5
0.55
0.6
0.65
0.7
Phase
Fig. 7. J ad H bad igh
ve f eg e y
b wih de ed e iida vaiai wih dahed
ie ad wih id ie e
day e
ie i
ded.
he di(cid:27)ee
e i he e bh wk (cid:28)d a igi(cid:28)
a
de
eae i a λ8190Å ie egh ve he ie fa
e f
he e
day a i eg a a
e e
e f iadiai
e(cid:27)e
. We eak ha if he iadiaed eegy de
eeae dee y i he ahee b ai y hea he
5. S ay
e ahei
aye i ay ed
e he vei
a e
ea e gadie ad heefe ead a de
eae i a
We de ed he e iida vaiai i JHKs igh
ve
abi ie egh wih a(cid:27)e
ig he
i
f eg eiae he ye aaee ad he
i
adiai aiig f deee ahei
aye. Thi i
b i f he e
day he a igh Tab e 2. Uig
i ie wih he iveigai f iadiai e(cid:27)e
CV
he
a ibaed (cid:29) xe we ifeed a e iva e b a
kbdy
e
daie by Baa a
hi d A ad 2002. Thei
e iiay dy hw ha f a yi
a CV iadiai
ead a igi(cid:28)
a
hage i he eea e
e
eea e f 3100± 500 K ad we deived a dia
e
he ye f 115 ± 30
.
Uig 3D e
ie aig e
hi e we deived he
f he e ahei
aye f he e
day a
fa
e bighe diib i f he eg a
ei di
.
whee he a ie
e f eavig he deee aye
Wih he aid f he ey f he e
ie a we (cid:28)d
he τ = 1 egi whee he R
i i d
ed
a di
ha f eig ag e f α = 2o
. The di
bighe
y a(cid:27)e
ed.
diib i hw a ayei
e a g he ga
di
eea e ae highe ha he ifeed
ea a je
y i he ie di
egi
e he whie
by F99. We (cid:28)d eea e f ∼ 3500K i a bad
whi e hey (cid:28)d eea e f ∼ 3000K i H . Thi i
i ie wih he fa
ha H bad igh
ve ha 2.1
dwaf i a bad R ∼ 0.1RL1 idi
aig he exie
e
f ga ea ve(cid:29)w i eg.
The a i de f he bia h de
eae wih i
y e (cid:29) x ha he 1994 Se igh
ve f F99 a
eaig wave egh. The h i de ed a a exeded
he efee
e hae φ = 0.25 whee he
ib i f
BS aii
eii h d be iia . daa ad
he f F99 wee
e
ed ee
ive y 5 ad 6 week afe
a b . The beved di(cid:27)ee
e i di
eea e
d be ex aied if he a
ei di
f eg w y
bigh
aed a he edge f he a
ei
∆θ ≃ 30o
di
wih bighe eea e f ∼ 10000K i J ad
H ad ∼ 6000K i he Ks bad. The hae f axi
f he bia h i R igh
ve φ = 0.74 de
i
ide wih he e i he i
a φ = 0.9 ee Fig. 1
dw d ig ie
e
e. A eaive y he beved
f Wd Cawfd 1986 de i
i
ide wih he
di(cid:27)ee
e i igh
ve (cid:29) x ad di
eea e
d
hae f axi eii f a adia y eiig bigh
be a
ed f by he 10%
eaiy i (cid:29) x
a ibai
f bh wk.
. The e
ie ii f he bigh di(cid:27)e f i
heei
a azi ha ii by ∼ 5o
ad i die
i f
axi eii ifeed f he igh
ve di(cid:27)e
S i . C. R e R.G.. 1998 RAS 299 768.
f he die
i f adia eii.
Se
ke D.W. Ei
k E.F. Whieb F.C. 1979 AS 41
by ∼ 54o
de a
h he deh ad widh f he e
day
501.
Skide W. a E. we S. B. Ciadi D. R. i efai S. .
e
ie a izeab e fa
i f he a
ei di
be
Dhi V. S. 2000 RAS 318 429.
a e. The ifeed eea e f he ie ad e
Sk kie .F. e a . 2006 A 131 1163.
heihee f he e
day a ae he ae wihi
Szkdy . ae . 1986 A 92 483.
he
eaiie idi
aig ha i iai e(cid:27)e
ae
Wae B. 1995 Caa
yi
Vaiab e Sa. Cabidge Uiv.
eg igib e f he R
i .
e Cabidge.
Wa C. A. Dhi V. S. R e R. G. . S
hwe A. D.
2003 RAS 341 129.
W f S. ae . .; e .; Bawig . Sheb R.;
6. A
kw edge
Baebae . 1993 AA 273 160.
Wd . . Cawfd C. S. 1986 RAS 222 645.
We hak D ad f iig he efee
e f he
Wd . . e . Beia G. Wade R. A. Dgh e D.
S R
Fak Gibbi ad Chi Be f ef i
Wae B. 1986 RAS 219 629.
Wd . . e . Beia G. Wade R. A. 1989 A 341
fai ab he W RCA ad he ay ef
974.
eee f ef
e whi
h he ed ive he e
eai f e . Thi wk wa aia y ed
by C Bazi h gh eea
h ga 62.0053/01
1 ADCT /i ei. T.R. a
kw edge (cid:28)a
ia
f C. R.B. a
kw edge (cid:28)a
ia
f C Bazi h gh ga 300.354/96 7 ad
200.942/2005 0. Thi b i
ai ake e f daa d
f he Tw i
A Sky S vey whi
h i a ji
je
f he Uiveiy f aa
h e ad he faed
eig ad Aa yi Cee/Ca ifia i e f
Te
h gy f ded by he aia Aea i
ad Sa
e
Adiiai ad he aia S
ie
e F dai.
hi eea
h we have ed ad a
kw edge wih hak
daa f he AAVS eaia Daa bae whi
h ae
baed bevai
e
ed by vaiab e a beve
w dwide.
Refee
e
Baia R. Seie . E. 1993 AA 277 331.
Baia R. a e R eda . a afi E.T. ah T.R.
Seegh D. 2005 AA 444 201.
Baa T. a
hi d . A ad F. 2002 i The hyi
f
Caa
yi
Vaiab e ad Re aed b je
AS Cfee
e
Seie V . 261 ed. B. T. Gaei
ke . Be ea .
Rei
h. . 49.
Beeka G. Se . ay T. e ie C. 2000 RAS
318 913.
Bee . S. Be . . 1988 AS 100 1134.
Bge B.W. Baia R. Caa .S. 2007 i eaai.
Caee . . 2001 A 121 2851.
C ae A. 1998 AA 335 647.
Davey S. Sih R.C. 1992 RAS 257 476
Diaz Cdvé . Giéez A. 1992 AA 259 227.
Fig C.S. Rbi E.. We h W.F. Wd . . 1999 A
523 399 F99.
ai T. E.; be . .; we S. B. 2005a A 129 2400.
ai T. E.; we S. B.; h . . 2005b B . A. A.
S
. 207 7016.
ghe S. . R
he . Dhi V. S. 1996 W RCA Ue
G ide v1.0 aa
ew G f Te e
e a a a
é . . Ga ia F. 2001 AA 366 359.
z R. . 1979 AS 40 1.
egge S.. 1992 AS 82 351.
i efai S.. Dhi V.S. ah T.R. a afi E. T. 2001
RAS 327 475.
ah T. R. 1988 RAS 231 1117.
ai .S. e D. . Sih R.C. 1987 RAS 224 1031.
eye F. eye feie E. 1982 AA 106 34.
a ebeg . 1972 A 175 417.
e W. . F aey B.. Te k ky S.A. Vee ig W.T. 1986
ei
a Re
ie Cabidge Uiveiy e
R e R.G.. va aadij . Tibege . 1992 AA 254 159.
Sai R. . Baia R. 2006 A 131 2185.
Saa . . 1989 AA 224 98.
Sak . . 1992 AUS 151 83.
|
astro-ph/0502051 | 1 | 0502 | 2005-02-02T21:07:34 | Deposing the Cool Corona of KPD 0005+5106 | [
"astro-ph"
] | The ROSAT PSPC pulse height spectrum of the peculiar He-rich hot white dwarf KPD 0005+5106 provided a great surprise when first analysed by Fleming, Werner & Barstow (1993). It defied the best non-LTE modelling attempts in terms of photospheric emission from He-dominated atmospheres including C, N and O and was instead interpreted as the first evidence for a coronal plasma around a white dwarf. We show here that a recent high resolution Chandra LETGS spectrum has more structure than expected from a thermal bremsstrahlung continuum and lacks the narrow lines of H-like and He-like C expected from a coronal plasma. Moreover, a coronal model requires a total luminosity more than two orders of magnitude larger than that of the star itself. Instead, the observed 20-80 AA flux is consistent with photospheric models containing trace amounts of heavier elements such as Fe. The soft X-ray flux is highly sensitive to the adopted metal abundance and provides a metal abundance diagnostic. The weak X-ray emission at 1 keV announced by O'Dwyer et al (2003) instead cannot arise from the photosphere and requires alternative explanations. We echo earlier speculation that such emission arises in a shocked wind. Despite the presence of UV-optical O VIII lines from transitions between levels n=7-10, no X-ray O VIII Ly alpha flux is detected. We show that O VIII Lyman photons can be trapped by resonant scattering within the emitting plasma and destroyed by photoelectric absorption. | astro-ph | astro-ph |
Accepted for publication in the Astrophysical Journal
Deposing the Cool Corona of KPD 0005+5106
Jeremy J. Drake1 and Klaus Werner2
1Smithsonian Astrophysical Observatory, MS-3,
60 Garden Street,
Cambridge, MA 02138
[email protected]
2Institut fur Astronomie und Astrophysik, Sand 1, 72076 Tubingen, Germany
[email protected]
ABSTRACT
The ROSAT PSPC pulse height spectrum of the peculiar He-rich hot white
dwarf KPD 0005+5106 provided a great surprise when first analysed by Flem-
ing, Werner & Barstow (1993). It defied the best non-LTE modelling attempts
in terms of photospheric emission from He-dominated atmospheres including
C, N and O and was instead interpreted as the first evidence for a coronal
plasma around a white dwarf. We show here that a recent high resolution
Chandra LETGS spectrum has more structure than expected from a thermal
bremsstrahlung continuum and lacks the narrow lines of H-like and He-like C
expected from a coronal plasma. Moreover, a coronal model requires a total
luminosity more than two orders of magnitude larger than that of the star it-
self. Instead, the observed 20-80 A flux is consistent with photospheric models
containing trace amounts of heavier elements such as Fe. The soft X-ray flux
is highly sensitive to the adopted metal abundance and provides a metal abun-
dance diagnostic. The weak X-ray emission at 1 keV announced by O'Dwyer
et al (2003) instead cannot arise from the photosphere and requires alternative
explanations. We echo earlier speculation that such emission arises in a shocked
wind. Despite the presence of UV-optical O VIII lines from transitions between
levels n = 7-10, no X-ray O VIII Lyα flux is detected. We show that O VIII Ly-
man photons can be trapped by resonant scattering within the emitting plasma
and destroyed by photoelectric absorption.
Subject headings: stars: activity -- stars: coronae -- stars: white dwarfs --
stars: mass loss -- X-rays: stars
-- 2 --
1.
Introduction
The hot 106-107 K coronae on the Sun and other late-type stars are believed to be sus-
tained by mechanical energy in their outer convection zones, which is dissipated at the surface
through the medium of magnetic fields generated and amplified by differential rotation and
convection in the interior. This paradigm is reinforced by the decline and disappearance of
X-ray emission that occurs toward hotter stars at spectral type A, which also corresponds
to the disappearance of outer convection zones. While the extensive Einstein stellar survey
(Vaiana et al. 1981) detected copious X-rays from earlier O- and B-type stars in addition
to later types, this has generally been attributed to emission from radiatively-driven super-
sonic shock-heated winds rather than coronal plasma (Cassinelli 1982), although the possible
contributions from magnetically-confined plasma have recently been reconsidered based on
recent Chandra and XMM-Newton high resolution spectra (e.g. Miller 2002).
In the above context, the low resolution ROSAT Position-Sensitive Proportional Counter
(PSPC) pulse height spectrum of the unusual hot He-rich white dwarf KPD 0005+5106
(Downes et al. 1985, WD 0005+511) presented a major surprise. Fleming et al. (1993)
found that they could not match this spectrum using the best available spectral models
computed using non-LTE model atmospheres containing He, C, N and O. Instead, they
found that a satisfactory fit to the data could be obtained from thermal bremsstrahlung and
optically-thin plasma models with temperatures of 2 − 3 × 105 K. The apparent impropriety
of photospheric spectral models and success of thermal plasma models provided evidence
that the X-rays originated not from deeper atmospheric layers but from a coronal plasma
encircling the star. KPD 0005+5106 then became the first white dwarf thought to have a
corona, albeit a cool one compared to the 106 − 107 K coronae of late-type stars. As Fleming
et al. (1993) noted, this result was quite novel and surprising, since all previously observed
X-ray emission from white dwarfs had been photospheric in nature. Moreover, while DA
and DB stars can possess significant convection zones corresponding to H and He ionization
layers and might conceivably sustain magnetic dynamo activity, this is not the case for the
much hotter 1.2 ×105 K effective temperature of KPD 0005+5106 (Werner et al. 1994) where
He is completely ionized.
A corona about KPD 0005+5106 is perhaps a less surprising proposition when viewed
in the light of other peculiarities of its spectrum. Emission lines of C V, N V and O VIII
have been detected at UV and optical wavelengths, prompting the suggestion of ongoing,
and possibly shock-heated, mass loss (Werner et al. 1994; Sion & Downes 1992; Sion et al.
1997). More recently, O'Dwyer et al. (2003) and (see also Chu et al. 2004) have reported a
detection of weak X-ray emission at 1 keV based on a re-analysis of deeper ROSAT PSPC
observations of KPD 0005+5106, while Otte et al. (2004) have presented the discovery of an
-- 3 --
O VI-emitting nebula associated with the object.
In this paper, however, we show that the Chandra Low Energy Transmission Grating
Spectrograph and High Resolution Camera Spectroscopic detector (LETG+HRC-S) spec-
trum of KPD 0005+5106 can be qualitatively modelled as photospheric emission, and does
not require a coronal explanation. The excess flux found by Fleming et al. (1993) in the 0.2-
0.3 keV range arises from deeper atmospheric layers revealed by a lower continuous opacity
in this region. Difficulties with earlier photospheric modelling attempts are shown to have
arisen because of the neglect of trace heavier elements such as Fe. The 1 keV emission found
by O'Dwyer et al. (2003) lies below our detection threshold, but cannot be explained by
photospheric models.
2. Observations and Analysis
KPD 0005+5106 was observed by Chandra using the LETG+HRC-S in its standard
configuration on 2000 October 25 between UT 05:16 and 10:58 for a total of 19114s, after
correction for instrument deadtime and bad event filtering.
Initial reduction of satellite telemetry was performed by the Chandra X-ray Center
Standard Data Processing software, but was completely reprocessed by us using CIAO 3.1.
The analysis described here is based on the Level 2 products of this reprocessing. The
scientific analysis was undertaken using the PINTofALE1 IDL2 software suite (Kashyap &
Drake 2000). The spectrum of KPD 0005+5106 was extracted by summing the events within
a window 3.6′′ in width in the cross-dispersion direction, and background was estimated in
two strips of width 18′′ each located on either side of the spectral trace. The final extracted
spectrum, binned at 5 A intervals, is illustrated in Figure 1. The distribution of the dispersed
photon events is consistent with a continuum source: no narrow emission line features were
discernible.
Our analysis involved comparison of the observed LETGS spectrum of KPD 0005+5106
with the emergent spectra computed for a range of line blanketed non-LTE model atmo-
spheres and for optically-thin collision-dominated coronal models. Coronal radiative loss
models were computed using PINTofALE, employing line and continuum emissivities com-
puted using atomic data from the CHIANTI database version 4.02 (Young et al. 2003) and
the ionization equilibrium of Mazzotta et al. (1998).
1Freely available from http://hea-www.harvard.edu/PINTofALE/
2Interactive Data Language, Research Systems Inc.
-- 4 --
Model atmosphere computations were performed using the PRO2 code (Werner et al.
2003) assuming hydrostatic and radiative equilibrium. Model atoms and atomic data used
for detailed non-LTE calculations have been described recently by Werner et al. (2004).
Constraints for the input model parameters are provided by earlier analyses of UV spec-
tra obtained by the HST Faint Object Spectrograph (FOS) and Goddard High Resolution
Spectrograph (GHRS), and optical spectra obtained at the 3.5 m telescope at Calar Alto
(Werner et al. 1994, 1996). Additionally, we have examined more recent Far Ultraviolet Spec-
troscopic Explorer (FUSE) spectra of KPD 0005+5106 to provide further constraints on the
abundances of the elements O and Fe that exhibit lines in the FUSE bandpass. From the
lack of Fe lines an upper abundance limit was determined by Miksa et al. (2002). From O VI
lines we have determined the O abundance with the models computed for the present paper.
Our adopted nominal parameters for KPD 0005+5106 are listed in Table 1. For a single test
calculation including sulfur we assume a solar S abundance (by mass fraction). The absolute
flux calibration for our models is provided by HST UV observations to high precision, and
for comparisons with Chandra spectra we normalise the models to the observed flux density
of 4.0 × 10−12 erg cm−2 s−1 A−1 at 1200 A.
The reasoning of Fleming et al. (1993) leading to the conclusion that KPD 0005+5106
has a corona was based on the comparison of low resolution X-ray pulse height spectra with
model predictions for atmospheres containing C, N and O. We argue in this paper that
difficulties in matching the X-ray data with photospheric models was primarily due to the
neglect of heavier elements that can contaminate the photosphere. While settling under the
strong gravity of the white dwarf tends to empty the atmosphere of heavier elements, the
atmospheres of hotter stars can be substantially enriched by radiative levitation (e.g. Chayer
et al. 1995; Schuh et al. 2002).
Model fluxes for atmospheres with an effective temperature Tef f = 120000 K, with and
without Fe, are compared in Figure 2. Models include full non-LTE treatment of C, O, Ne, S
and Fe, whose abundances were adopted from Table 1. Also illustrated is the blackbody flux
density distribution for the same temperature. It is interesting to note that the blackbody
has much less short wavelength flux than do the more realistic models: the short wavelength
extension of the spectra of the latter arises because of changes in the opacity with wavelength
that expose the deeper, hotter atmospheric layers at these wavelengths. Addition of Fe with
an abundance equal to the best current upper limit (Table 1) however, dramatically reduces
the soft X-ray flux: clearly any photospheric soft X-ray emission from KPD 0005+5106 and
similar stars is critically dependent on the photospheric abundances of heavier elements, as
was also demonstrated earlier in the case of EUV fluxes at longer wavelengths for slightly
cooler stars (e.g. Barstow et al. 1995).
-- 5 --
Coronal and photospheric model spectra are compared to the observations in Figure 3.
Coronal models computed for the He-rich composition listed in Table 1 are dominated by
the He bound-free and free-free continua, and by the lines of He-like C near 40 A. While it
is tempting to interpret the peak in the observed flux between 40 and 45 A as being due to
these C V lines, there are no such lines visible in the spectrum where the observed counts are
more smoothly distributed. The absence of C V lines does not necessarily argue against the
coronal interpretation of the spectrum, however, since one might expect heavier elements to
gravitationally settle out of such a plasma. We therefore also computed models in which the
metal abundances were reduced by a factor of 100.
We confirm the result of Fleming et al. (1993) that the best-fit coronal spectrum has
a temperature in the range 2-3 × 105 K; our best-fit model corresponds to T = 2 × 105 K,
with an interstellar absorption component represented by a column of neutral hydrogen of
3.5 × 1020 cm−2. Qualitative comparison of the best-fit coronal model and data suggest,
however, that the latter have a more complex shape than can be achieved by the former.
Photospheric models both with and without Fe were computed, as was a test model
including S; a small subsection of the models investigated are illustrated in the lower panel
of Figure 3. It is clear that models with only C, O and Ne greatly overpredict the observed
X-ray flux by factors of 10 or more. Based on Figure 2, we see that the emergent soft X-
ray flux can be reduced by addition of Fe and the opacity this element provides at these
wavelengths. This is borne out in practice, and the photospheric spectrum that was found
to best match the observations corresponds roughly to the parameters Tef f = 120000 K and
Fe/He= −5.5, with a neutral hydrogen column density of 3 × 1020 cm−2. It is clear from
Figure 3 that there is no obvious reason to prefer coronal models over photospheric ones.
While we have not succeeded in producing a photospheric model that perfectly matches
the observations, we have achieved a good qualitative match by adding only plausible
amounts of Fe to the C, O and Ne abundance mixture. Trace amounts of other metals
such as Na, Mg, S, Ar, Ca and Ni could also affect significantly the soft X-ray spectrum, and
we could doubtless obtain a better match through their addition and subsequent optimisa-
tion of other parameters such as effective temperature, surface gravity and ISM absorption.
The test model including S, for example, introduced an absorption edge near 43 A with little
effect on the emergent flux at longer wavelengths; careful adjustment of the S abundance
could aid in a better prediction of the observed drop in flux between 40 and 45 A. Further
progress in this direction would demand some additional constraints on the abundances of
these other metals in order to limit the available parameter space to tractible proportions.
-- 6 --
3. Discussion and Conclusions
3.1. Photospheric vs Coronal Emission
We have shown that the Chandra LETG+HRC-S spectrum of KPD 0005+5106 in the
25-90 A range can be qualitatively modelled by photospheric emission from an atmosphere
containing He, C, O and Ne in expected amounts, with the addition of Fe at a level consistent
with the current observational upper limit. With these quite plausible models, there is no
requirement in the data to resort to more exotic solutions such as a coronal model. The
quality of the current short LETGS exposure does not allow us to rule out a significant
amount of coronal emission with a high degree of confidence, though the best fit model
based on purely coronal emission is statistically unacceptable with a reduced χ2 = 2.1.
We can, however, completely discount a coronal interpretation of the ROSAT and Chan-
dra spectra on energetic grounds. A coronal plasma model at a temperature of 2-3 × 105 K
absorbed by an ISM column of NH ∼ 1020 cm−2 and normalised to the observed ROSAT
spectrum has a total luminosity 2-3 orders of magnitude greater than than the star itself.
This is qualitatively illustrated in Figure 2, where the coronal model is shown normalised
relative to the photospheric models to the approximate relative level required to explain the
observed 20-80 A X-ray flux.
3.2. Remaining Puzzles
Despite the ability of photospheric models to explain the Chandra LETGS spectrum
of KPD 0005+5106 and the energetic impossibility of a coronal explanation, there are two
additional puzzles regarding the spectrum of this object at both shorter and longer wave-
lengths that cannot be explained by our photospheric models. These are the detection of
flux at ∼ 1 keV (12.4 A) by O'Dwyer et al. (2003), and the observation of ∆n = 1 emission
lines of O VIII between levels n = 7-10 at UV and optical wavelengths (Werner et al. 1994;
Sion & Downes 1992; Sion et al. 1997). It seems likely that the two are related.
We have verified that there is no trace of any flux near 12 A in the LETGS spectrum,
although the flux density reported by O'Dwyer et al. (2003) based on ROSAT observations (5
count s−1 in the Boron filter) lies below our LETGS detection limit in this short observation:
only ∼ 18 LETG+HRC-S counts are expected, spread over a fairly large detector area over
which the background is an order of magnitude larger.
More puzzling is the lack of O VIII lines in the Chandra data: the excited higher n states
will decay through cascades to lower n, and we would expect to see some fraction of this
-- 7 --
decay channel in the Lyman series resonance lines. Based on the Goddard High Resolution
Spectrograph UV spectrum of KPD 0005+5106 obtained on 1994 June 1 and analysed by
Werner et al. (1996), the O VIII 2976.57 A n = 8-7 transition flux is 0.07 ph cm−2 s−1. The
upper limit to the flux of the O VIII 18.98 A n = 2-1 doublet in the Chandra spectrum is
more than three orders of magnitude less than this.
There are two possible explanations for the absence of the O VIII Lyα lines: (1) the
source is variable and during the Chandra observation the O VIII lines were much weaker;
(2) the X-ray lines are suppressed by some mechanism. While the lack of any simultaneous
observations accompanying the Chandra pointing prevents us drawing definitive conclusions
regarding (1), we note that optical spectra of KPD 0005+5106 similar to those described
by Werner et al. (1994) and obtained at the Calar Alto 3.5m telescope in 1991 July, 1992
September and 1994 May, exhibit identical O VIII n = 10-9 and 9-8 lines and conclude that
large amplitude variability in O VIII is unlikely. Instead, the most likely explanation for the
lack of prominent X-ray O VIII lines is that they are suppressed.
We speculate that the O VIII Lyα lines are formed in a low density medium or wind
surrounding the white dwarf, as has been suggested by earlier workers (Sion & Downes 1992;
Sion et al. 1997; Werner et al. 1994).
In this case, the low n resonance lines might be
sufficiently optically thick to resonance scattering that line photons are destroyed by pho-
toelectric absorption as they undergo multiple scattering events within the emitting region.
The line centre optical depth, τ , in a He-dominated low-density medium can be written (e.g.,
Acton 1978; Mariska 1992):
τ = 1.16 · 10−14 ·
ni
nel
AZ
nH
ne
λfrM
T
neℓ
(1)
for ion fraction (ni/nel, element abundance AZ, oscillator strength f , temperature T , electron
density ne, atomic weight M, and where nHe/ne ∼ 0.5, and ℓ is the total path length along
the line of sight through the emitting plasma.
Taking O VIII 18.97 A 2p → 1s as an example, we have estimated the optical depth for
two limiting cases of the morphology of the emitting plasma: (a) a "corona" in hydrostatic
equilibrium, for which the scale height, or typical pathlength, ℓ, is much smaller than the
stellar radius, ℓ << R⋆; (b) an extended wind for which ℓ >> R⋆. We have used PINTofALE
(Kashyap & Drake 2000) to estimate the total volume emission measure, n2
eV , of the O VIII
emitting plasma based on the total of 25 ROSAT (Boron filter) counts observed near 1 keV in
5 ks by O'Dwyer et al. (2003). We assumed that the plasma is isothermal with T = 2 ×106 K
(corresponding to the peak in the O7+ ion fraction, such that ni/nel ∼ 1), is collision-
dominated, has a He-dominated composition as listed in Table 1 and is absorbed by an
ISM column density represented by a neutral H column of NH = 5 × 1020 cm−2. For the
-- 8 --
approximate distance of 270 pc estimated by Werner et al. (2004), the observed ROSAT
counts imply a volume emission measure n2
eV ∼ 1054 cm−3.
Adopting the stellar parameters of KPD 0005+5106 from Werner et al. (1994), the
pressure scale height for a He plasma in hydrostatic equilibrium, kT /mHeg⋆ = 6.2 × 106 cm,
or 0.2 % of the stellar radius and the plasma density is ∼ 1014 cm−3. Based on Equation 1,
the optical depth in O VIII Lyα is τ = 17 -- ie extremely large. For a much more extended
plasma, the line centre optical depth is τ ∼ 200 for ℓ ∼ 10R⋆ and remains very large (τ > 10)
for emitting size scales up to 104R⋆.
The large optical depths in the O VIII Lyα resonance line imply that these line photons
are trapped within the plasma and must undergo many scattering events before escaping.
If the plasma has a temperature of ∼ 2 × 106 K, it will contain a significant fraction of
H-like and He-like ions among the light elements C, N and O, together with ions with
n = 3 ground states of elements such as Fe if these are present. The total photoelectric
absorption cross-section for our postulated plasma, computed using the Hartree-Dirac-Slater
photoionisation cross-sections of Verner et al. (1993) and Verner & Yakovlev (1995), together
with ion fractions from Mazzotta et al. (1998), amounts to σ ∼ 10−23 cm2. For the plasma
radial extents considered above, for which O VIII Lyα is optically thick, the typical column
density of the plasma in terms of He nuclei ranges from several 1020 to 1022 cm−2. Absorption
for weaker, optically thin lines and continuum is therefore negligible. However, the trapped
O VIII line photons must undergo such a large number of scattering events that they are
destroyed by photoelectric absorption before they can escape. Similar arguments apply to
higher Lyman series lines; while their f-values are smaller, they are also effectively quenched
through decays to levels n > 1 and then eventually through Lyα as soon as scattering depths
become significant. O VIII 16.01 A Lyβ, for example, is lost to resonant scattering in decays
through the n = 3-2 Balmer line at 102.43 A, and then through Lyα. Unfortunately, the
ISM absorbing column to KPD 0005+5106 is too large for the Balmer lines to be visible,
though a future detection of O VIII Baα in other similar objects to KPD 0005+5106 would
provide a useful test of this model.
If the ∼ 1 keV emission found by O'Dwyer et al. (2003) is indeed due to a wind or
outflow, then an optically-thin plasma model with a temperature of 2 × 106 K accounting for
the observed ROSAT counts at this energy has a total luminosity of ∼ 5 × 1031 erg s−1, or
10−4 times the stellar bolometric luminosity. The 0.1-2.5 keV X-ray to bolometric luminosity
ratio is ∼ 2 × 10−5 -- a somewhat higher value than observed for OB stars, whose X-rays are
believed to originate in shocked winds and for which the X-ray to bolometric luminosity
ratio LX /Lbol ∼ 10−6-10−8 (e.g. Berghoefer et al. 1997).
-- 9 --
4. Conclusions
We have shown that a coronal plasma is unable to match the Chandra LETG+HRC-S
20-80 A spectrum of KPD 0005+5106 and that coronal models are energetically implausible
as an origin for this observed soft X-ray flux. This part of the soft X-ray spectrum of
KPD 0005+5106 can instead be explained by photospheric models containing trace amounts
of heavier elements. Photospheric models do not, however, explain the soft X-ray emission at
shorter wavelengths (∼ 12 A; 1 keV) revealed recently by O'Dwyer et al. (2003) and discussed
in more detail by Chu et al. (2004). The origin of this emission remains mysterious, though
an outflow or wind with a temperature of ∼ 2 × 106 K is able to explain both the X-rays
and the presence of high n O VIII UV-optical emission lines but an absence of significant
O VIII Lyα flux in the Chandra spectrum. In this scenario, the Lyα photons are trapped by
resonance scattering and destroyed by photoelectric absorption. If an outflow is responsible
for the 1 keV X-rays, the total luminosity of this plasma amounts to ∼ 10−4Lbol, which is a
factor of 100 or so larger than for winds of OB stars.
We thank the NASA AISRP for providing financial assistance for the development of
the PINTofALE package, and the CHIANTI project for making publicly available the results
of their substantial effort in assembling atomic data useful for coronal plasma analysis. Brad
Wargelin is thanked for useful comments. JJD was supported by NASA contract NAS8-
39073 to the Chandra X-ray Center during the course of this research. X-ray data analysis
in Tubingen is supported by the DLR under grant 50 OR 0201.
Acton, L. W. 1978, ApJ, 225, 1069
REFERENCES
Barstow, M. A., Holberg, J. B., Werner, K., & Nousek, J. A. 1995, Advances in Space
Research, 16, 73
Berghoefer, T. W., Schmitt, J. H. M. M., Danner, R., & Cassinelli, J. P. 1997, A&A, 322,
167
Cassinelli, J. P. 1982, Advances in Space Research, 2, 67
Chayer, P., Fontaine, G., & Wesemael, F. 1995, ApJS, 99, 189
Chu, Y., Gruendl, R. A., Williams, R. M., Gull, T. R., & Werner, K. 2004, AJ, 128, 2357
Downes, R. A., Liebert, J., & Margon, B. 1985, ApJ, 290, 321
-- 10 --
Fleming, T. A., Werner, K., & Barstow, M. A. 1993, ApJ, 416, L79+
Kashyap, V. L. & Drake, J. J. 2000, Bulletin of the American Astronomical Society, 32, 1227
Mariska, J. T. 1992, The solar transition region (Cambridge Astrophysics Series, New York:
Cambridge University Press, -- c1992)
Mazzotta, P., Mazzitelli, G., Colafrancesco, S., & Vittorio, N. 1998, A&AS, 133, 403
Miksa, S., Deetjen, J. L., Dreizler, S., Kruk, J. W., Rauch, T., & Werner, K. 2002, A&A,
389, 953
Miller, N. A. 2002, in ASP Conf. Ser. 277: Stellar Coronae in the Chandra and XMM-
NEWTON Era, ed. F. Favata & J. J. Drake, San Francisco, 379
O'Dwyer, I. J., Chu, Y., Gruendl, R. A., Guerrero, M. A., & Webbink, R. F. 2003, AJ, 125,
2239
Otte, B., Dixon, W. V. D., & Sankrit, R. 2004, ApJ, 606, L143
Schuh, S. L., Dreizler, S., & Wolff, B. 2002, A&A, 382, 164
Sion, E. M. & Downes, R. A. 1992, ApJ, 396, L79
Sion, E. M., Holberg, J. B., Barstow, M. A., & Scheible, M. P. 1997, AJ, 113, 364
Vaiana, G. S., Cassinelli, J. P., Fabbiano, G., Giacconi, R., Golub, L., Gorenstein, P., Haisch,
B. M., Harnden, F. R., Johnson, H. M., Linsky, J. L., Maxson, C. W., Mewe, R.,
Rosner, R., Seward, F., Topka, K., & Zwaan, C. 1981, ApJ, 245, 163
Verner, D. A. & Yakovlev, D. G. 1995, A&AS, 109, 125
Verner, D. A., Yakovlev, D. G., Band, I. M., & Trzhaskovskaya, M. B. 1993, Atomic Data
and Nuclear Data Tables, 55, 233
Werner, K., Deetjen, J. L., Dreizler, S., Nagel, T., Rauch, T., & Schuh, S. L. 2003, in ASP
Conf. Ser. 288: Stellar Atmosphere Modeling, 31 -- 1
Werner, K., Dreizler, S., Heber, U., Rauch, T., Fleming, T. A., Sion, E. M., & Vauclair, G.
1996, A&A, 307, 860
Werner, K., Heber, U., & Fleming, T. 1994, A&A, 284, 907
Werner, K., Rauch, T., Barstow, M. A., & Kruk, J. W. 2004, A&A, 421, 1169
-- 11 --
Young, P. R., Del Zanna, G., Landi, E., Dere, K. P., Mason, H. E., & Landini, M. 2003,
ApJS, 144, 135
This preprint was prepared with the AAS LATEX macros v5.2.
-- 12 --
Fig. 1. -- The Chandra LETG+HRC-S spectrum of KPD 0005+5106; shaded regions rep-
resent 1σ uncertainties. No evidence for narrow lines is present in the data. We therefore
show here the spectrum binned at 5 A intervals in order to illustrate the observed continuum
structure.
-- 13 --
Fig. 2. -- Emergent spectra in the UV-X-ray range for non-LTE model atmospheres com-
puting using the PRO2 code (Werner et al. 2003) for the parameters and abundances listed
in Table 1. Dashed curves illustrate the spectra absorbed by an intervening ISM column of
4 × 1020 cm−2. The model including Fe assumed Fe/He= −5, consistent with the current
upper limit based on UV and optical spectra. Also shown is a coronal model computed for a
temperature T = 2 × 105 K, normalised to the relative level required to explain the ROSAT
observations presented by Fleming et al. (1993).
-- 14 --
Fig. 3. -- Comparison of the observed LETGS spectra (shaded region) with different coronal
(upper) and non-LTE photospheric model (lower) predictions. Coronal models are shown
both at the theoretical resolution of the LETGS, and binned at the same 5 A intervals as
the observations. Models are attenuated by an ISM absorbing column of NH = 4 × 1020.
Also shown in the lower panel is the "best fit" model ("120kK CONe+Fe/3") for which
NH = 3 × 1020.
-- 15 --
Table 1: Fundamental Parameters and Abundances for KPD 0005+5106
Teff
log g
H/He
C/Hea
O/He
Ne/He
S/He
Fe/He
NH
F1200
b
120000 ± 10000 K
7.0 ± 0.5
≤ −0.7
−3 ± 0.5
−3.8 ± 0.5
≤ −8
−4.4
≤ −5
(5 ± 1.1) × 1020 cm−2
4.0 × 10−12
aX/He= log10 n(X)/n(He)
bFlux density at 1200 A in units of erg cm−2 s−1 A−1
|
astro-ph/0402558 | 2 | 0402 | 2004-02-25T07:23:24 | Properties of the linearly polarized radiation from PSR B0950+08 | [
"astro-ph"
] | Measurements of average pulse profiles made with a single linear polarization over the range 41-112 MHz are presented for PSR B0950+08. We show that the observed variable structure of the pulse profiles is a result of Faraday sinusoidal modulation of the pulse intensity with frequency. The rotation measure corresponding to this effect, RM = 4 rad/m^2, is about 3 times greater than the tabulated value of RM = 1.35 rad/m^2. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. p0950
(DOI: will be inserted by hand later)
November 1, 2018
4
0
0
2
b
e
F
5
2
2
v
8
5
5
2
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Properties of the linearly polarized radiation from PSR
B0950+08
T. V. Shabanova and Yu. P. Shitov
Astro Space Center, P. N. Lebedev Physical Institute, Leninskij Prospect 53, 117924 Moscow, Russia
e-mail: [email protected], [email protected]
Received 30 July 2002 / Accepted 26 November 2003
Abstract. Measurements of average pulse profiles made with a single linear polarization over the range 41 -- 112
MHz are presented for PSR B0950+08. We show that the observed variable structure of the pulse profiles is a
result of Faraday sinusoidal modulation of the pulse intensity with frequency. The rotation measure corresponding
to this effect, RM ≈ 4 rad m−2, is about 3 times greater than the published value of RM = 1.35 rad m−2 (Taylor
et al. 1993).
Key words. stars: neutron -- pulsars: general -- pulsars: individual: PSR B0950+08
1. Introduction
PSR B0950+08 is well-studied over a wide frequency range from 24 to 10500 MHz. It exhibits a single pulse profile
at frequencies above 400 MHz and a double pulse profile at low frequencies in the range 24 -- 112 MHz (Hankins et al.
1991; Kuzmin et al. 1998). There is an interpulse occurring approximately 152◦ ahead of the main pulse and a bridge
of emission between the interpulse and main pulse (Lyne & Rickett 1968).
The study of average pulse profiles from a large number of pulsars has shown that a pulse profile obtained by
averaging several hundred individual pulses is very stable for most pulsars (Helfand et al. 1975; Rathnasree & Rankin
1995). Profile changes can be caused by such phenomena as mode changing and nulling which occur in a small fraction
of pulsars. The narrowband profile changes which are observed for the pulsar B0950+08 in the range 41 -- 112 MHz
and discussed in this paper are consistent with the effects of Faraday rotation on a linearly polarized pulse emission
received by a linearly polarized antenna. Pulsars with considerable linear polarization observed with a single linear
polarization show frequency-dependent profiles because of Faraday rotation of the plane of polarization in the inter-
stellar medium. Pulse profiles obtained for different epochs show time-dependent profile shapes caused by a variation
of the electron density or magnetic field in the propagation path, mainly due to varying ionospheric contribution to
the rotation measure. At the Pushchino Radio Astronomy Observatory (PRAO), the effect of Faraday modulation of
the pulse intensity with frequency at the output of the multi-channel receiver is used for the measurements of linear
polarization characteristics of pulsars and the estimation of their rotation measure (Vitkevich & Shitov 1970; Shitov
1971; Suleymanova 1989).
Changes of the average pulse profile at meter wavelengths for the pulsar B0950+08 were first noticed by Smirnova &
Shabanova (1992). They found that the pulse profile is variable in time and shows narrowband changes with frequency.
The explanation of the observed phenomenon by Faraday rotation was problematic because the predicted effect of
the interstellar Faraday rotation for the tabulated value of RM = 1.35 rad m−2 (Taylor et al. 1993) was negligible
across the receiver bandpass at 102 MHz. The authors supposed that the narrowband changes of the pulse profile may
be due either to the pulsar's intrinsic narrowband emission, manifesting only at low frequencies, or to scintillations
of spatially separate sources. However, the low time resolution of their observations (about 10◦ of longitude) made a
detailed analysis of this phenomenon impracticable.
The main purpose of this paper is to explain the observed narrowband profile changes of PSR B0950+08. The
techniques used for observations over the frequency range 41 -- 112 MHz are described in §2. In §3 a large set of average
profiles is investigated with respect to the shape changes with time. In §4 we study the narrowband changes of the
pulse profile across the 2.56-MHz bandpass and establish a relation between the shape changes with time and the
Send offprint requests to: T. V. Shabanova
2
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
Table 1. Parameters of the DKR observations
Parameter
Observing frequency (MHz)
Receiver time constant (ms)
Sampling interval (ms)
One bandwidth (MHz)
Dispersion pulse broadening (ms)
Total bandwidth (MHz)
Dispersion time delay (ms)
Number of pulses averaged
Value
88.57
1
0.8192
0.02
0.7
2.56
91
3770
62.15
1
0.8192
0.02
2
2.56
263
2200
41.07
3
2.9696
0.02
7
2.56
911
2160
shape changes with frequency. §5 discusses the frequency dependence of the pulse profile after removing the effect
caused by interstellar scintillation. In §6 we compute Faraday rotation effects using a numerical polarization model of
the pulse profile. §7 describes the narrowband changes of the average profile at the lower frequencies of 88, 62 and 41
MHz. §8 presents the results of the timing data analysis for the pulsar. In §9 we conclude that the profile changes are
due to the Faraday rotation effect.
2. Observations
Observations with a high time resolution of average pulse profiles and individual pulses from the pulsar B0950+08
were carried out at PRAO in three observing sessions from April 1996 to May 2002.
The first session took place from April 1996 through July 2001 and included regular timing observations of the
pulsar with the BSA radiotelescope which operated at 102.7 MHz until May 1998, and at 111.3 MHz since November
1998. The BSA large phased array radiotelescope, making up a linearly polarized transit antenna with 30000 m2
effective area and a beam size of about (3.5 / cos δ) minute, provides 3.3 minutes duration observations at the pulsar
declination δ ≈ 8◦. A 32 channel × 5 kHz spectrometer covering a total bandwidth of 160 kHz was used to provide
high frequency resolution. The receiver time constant was 0.3 ms. The data were sampled at intervals of 0.2048 ms. An
average pulse profile in each channel was formed by synchronous averaging of 770 individual periods with a predicted
topocentric pulsar period. After dispersion removal all the channel profiles were summed to form an averaged pulse
profile for a single observation. A narrow bandwidth of one 5-kHz channel limited the pulse broadening from the
interstellar dispersion to 0.09 ms. The observation window width was 200◦ of longitude. A total of 264 average pulse
profiles were derived at two observing frequencies around 102.5 and 112 MHz for a 5-yr span of observations.
The second session included observations of individual pulses and was made using the BSA telescope at an observing
frequency of 111.87 MHz during 2001 and 2002. The 128 spectral channels, each of bandwidth 20 kHz, were used to
record individual pulses for 3.3 minutes of the BSA transit time. The receiver time constant was 1 ms and the sampling
interval was 0.512 or 0.8192 ms. The signal dispersion through one 20-kHz channel causes pulse broadening of 0.35
ms.
The third session was carried out using the East-West arm of the DKR-1000 radiotelescope at three lower frequencies
of 88.57, 62.15 and 41.07 MHz in late 2001 and early 2002. Individual pulses were recorded during 16 minutes of the
DKR transit observation time with the 128 channel × 20 kHz spectral radiometer. The parameters of the DKR
observations are listed in Table 1. The effects of the interstellar dispersion in the total bandwidth of 2.56 MHz at
three observing frequencies were 91, 263 and 911 ms, respectively. The dispersion time delay at 62.15 and 41.07 MHz
was more than the 253-ms pulsar period. To compensate for this time delay, the signal was recorded blocks of 5 full
periods at 62.15 MHz and blocks of 12 full periods at 41.07 MHz. From each block, the signal was only analyzed for
either the 3 or 8 first successive periods at each observing frequency, respectively. The average profile was formed by
addition of 2200 pulse periods at 62.15 MHz and 2160 pulse periods at 41.07 MHz.
The pulsar B0950+08 shows a high degree of linear polarization of about 90% and a smooth change in the position
angle through the interpulse and main pulse, as measured at 150 and 151 MHz (Schwarz & Morris 1971; Lyne et al.
1971). Both the BSA and DKR radiotelescopes receive only one polarization. For our observations, the published value
of a rotation measure of RM = 1.35 rad m−2 (Taylor et al. 1993) predicts negligible effect of Faraday rotation across
the 2.56-MHz bandpass. The expected periods of Faraday modulation of the pulsar emission with frequency are 18
MHz at the observing frequency of 112 MHz, 14 MHz at 102.7 MHz, 9 MHz at 88.6 MHz, 3 MHz at 62 MHz and 0.9
MHz at 41 MHz. This effect produces a modulation of the pulse emission spectrum of less than 15% at the frequencies
within the 102 -- 112 MHz range and is significant only at two lower frequencies of 62 and 41 MHz. The influence of a
varying ionosphere on our measurements for different days is also small. Most observations of the pulsar were made
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
3
at night time or close to it (during autumn or winter), when an ionospheric contribution to the rotation measure RM
is estimated to be less than 0.4 rad m−2 for the PRAO site (Udaltsov & Zlobin 1974).
3. The changes of the average profile with time in the range 102 -- 112 MHz
We investigated average profiles using timing observations over the interval 1996 -- 2001. The shape of the observed
pulse profiles strongly varies from one observation to another that we can see both double and triple profiles. To reveal
some regularities in the occurrence of the profiles with different shapes, we sorted available average profiles into groups,
each containing similar profile shapes. Each of the 264 observed profiles could be put into one of the profile groups. A
detailed analysis of these data showed that the shapes from different groups make up a set smoothly varying between
the main states of the double profile. Such a set of continuously varying shapes is presented in Fig. 1.
The shape (a) represents the main pulse, which has two well-resolved components with a separation of 10◦. The
shapes from (b) to (g) show the well-resolved third component, occurring 34◦ ahead of the main pulse and visible at
a longitude of about 170◦. Note that the unresolved component ahead of the main pulse in the range 151 -- 430 MHz
was observed earlier (Lyne et al. 1971; Hankins & Cordes 1981). The shape changes affect the whole profile and both
the components of the main pulse alter their amplitude, either amplifying or weakening alternately. Variations in the
relative amplitudes of all three components exhibit a periodic character.
The two profiles (c) and (d) of Fig. 1 have a lower signal-to-noise ratio (S/N) and show unusual shapes. These profiles
are similar to the adjacent profiles (b) and (f) having a better S/N ratio. Shapes (c) and (d) are of great importance to
confirm the periodic character of changes in the pulse profile. These shapes provide a transition between the first and
second dominant components of the main pulse. The state (c) where the peak of the second component is significantly
higher than the peak of the first one quickly changes to another state (d) where the peak of the first component
dominates.
4. Changes of pulse shapes with frequency
Individual pulses from the pulsar were observed at 111.87 MHz with the 128 channel × 20 kHz spectral radiometer.
The high quality data set was analyzed to investigate the frequency structure of the average profiles within bands of
different widths of 640, 160 and 20 kHz.
The pulse profiles were obtained by averaging 770 individual periods for each 32-channel quadrant of the filter
bank. The sequences of 4 resulting profiles are presented in Fig. 2 for three different observations. Detailed inspection
of each panel shows that the shape of the average profile changes markedly over the bandpass of 2.56 MHz. It is also
seen that the sequences shown are not identical although the same shapes occur in all the cases. An analysis of the
available data showed that the profile shapes vary with frequency in a strictly defined order and that the frequency
changes are a result of some modulation process having a phase drift with time. If the process observed is related to
Faraday rotation then the tabulated value of RM would be expected to be wrong. Therefore the predicted Faraday
effect cannot be negligible. Variable ionospheric Faraday rotation does not affect the profiles during the 3 minutes of
observation time but causes a drift of the frequency picture with time.
The receiver bandwidth is less than the frequency interval of the modulation process and so we observe only some
stages of this process on different days. Nevertheless, a period of the profile frequency changes can be estimated if we
analyze the records together, which contain overlapping shapes as shown in Fig. 2. Here the last shape at 109.95 MHz
of the panel (a) is similar to the first shape at 111.87 MHz of panel (b). This is also the case for panels (b) and (c). A
set of 12 plotted profiles covers the full cycle of the shape changes and shows how the frequency picture develops from
higher to lower frequencies, regardless of the particular value of the observing frequency. The transition between the
first and second dominant components, which was discussed in the preceding section, is clearly observed in panel (a)
between the shapes at 111.23 and 110.59 MHz and panel (c) between the shapes at 111.87 and 111.23 MHz. This
state indicates a transition to a new cycle of profile changes with frequency. The interval between these states is 6 -- 7
640-kHz bands, i.e. it is approximately equal to 4 MHz. Hence, having a 2.56-MHz band, we can nevertheless estimate
the frequency interval of the shape changes which can be several times wider than a bandwidth.
For a more detailed analysis of the profile frequency structure, individual pulses were averaged within narrower
160-kHz bands. Two sets of 16 resulting profiles are presented in Fig. 3 for two different epochs. They give a finer
frequency structure of the 640-kHz profiles, already displayed in Fig. 2a and b. In order to see the profile frequency
dependence inside one cycle of changes, the right hand sequence should be considered as an extension of the left hand
one. The first component of the profile at 111.87 MHz (panel (a)) weakens gradually toward lower frequencies and
becomes unresolved at 110.59 MHz. Beginning from the profile at 110.43 MHz, the first component dominates. The
last 4 profiles from 109.95 to 109.47 MHz of this sequence are similar to the first 4 profiles of the right hand sequence.
The second component becomes visible in the profile at 111.23 MHz (panel (b)). At 110.27 MHz, the two components
become comparable in amplitude. Then, the first component becomes again weaker and the picture is repeated. The
4
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
profile at 109.79 MHz of the panel (b) is similar to the first profile at 111.87 MHz of the panel (a). It should be noted
that the left and right sequences show different rates of shape changes with frequency. It may be caused by a varying
ionospheric contribution to the rotation measure on different days.
We see that the set of the profile shapes plotted in Fig. 1 is similar to the set of the shapes presented in Fig. 3.
This suggests that the profile changes with time are a result of the profile changes with frequency. When the receiver
bandpass is narrower than the frequency interval of the shape changes, the profile shape obtained will depend on
the location of the band inside this interval. Note that the pulse profiles of Fig. 1 were plotted in such a way that
the sequence of the profile shapes placed from top to bottom would correspond to the sequence of the profile shapes
observed within the receiver bandpass in the direction from a higher frequency to a lower one.
An analysis showed that the shape changes are clearly seen even in individual channels of 20-kHz width and these
changes cannot be caused by instrumental effects. The transition state may occupy a bandwidth from 40 to 160 kHz
for different observations. Weakening of the signal is a typical feature for all the observations where this state occurs.
5. Interstellar scintillations and the frequency structure of the profile changes at 112 MHz
It is well known that the pulse emission of PSR B0950+08 is modulated by interstellar scintillations. In order to
determine the type of frequency dependence of the profile changes we must exclude modulation due to scintillations.
The emission spectrum of the pulsar within the receiver bandpass was obtained using the following procedure. For
each 20-kHz channel out of 128, the area under the average profile was calculated on the longitude interval above the
0.1 level of the profile peak and normalized to the rms of noise outside the profile in the same channel. The spectrum of
scintillations is characterized by modulation of the area under the average profile with frequency and is estimated by
the width of an autocorrelation function ACF . For our observations at 112 MHz, modulation of the emission spectrum
of the pulsar is close to 80%. The width of the decorrelation band of the spectrum of scintillations defined by the
half-width at the (1/e) level of ACF is equal to 250 kHz. This agrees well with earlier measurements at 105 MHz
made by Shitov (1972).
The emission spectrum of the pulsar in the receiver bandpass was calculated by dividing the amplitude of the
average profile for each channel by the area under the average profile in this channel. Some examples of the observed
emission spectra at the longitudes of pulse peaks of the three components are given in the left panel of Fig. 4 for
different observations.
As is seen in this plot, the spectra of the two main components of the pulse profile intersect at two points, which
correspond to two basic shapes of the average profile. Point 1 corresponds to a transition state between the first and
the second dominant components of the main pulse (see triple profiles (c) and (d) in Fig. 1). Here the pulse peak of the
first component becomes higher than the pulse peak of the second component. Near this point, the pulse peak of the
third component ahead of the main pulse has the greatest value. Point 2 corresponds to a state when the pulse peaks
of both the dominant components have comparable amplitudes, and the amplitude of the third component approaches
its smallest value (see double profile (a) in Fig. 1).
Uniform behavior of the spectral curves for all the observations can specify that the narrowband changes in the
pulse shape are deterministic. A detailed analysis of these data showed that the frequency structure of the pulse profile
can be modelled by a sinusoid. We approximated the obtained spectra of the pulse emission by three sinusoidal curves,
defined for each component as
y1(k) = 0.20 sin (
y2(k) = 0.25 sin (
y3(k) = 0.19 sin (
2π
T
2π
T
2π
T
∗ k + φ) + 0.60 ,
∗ k − 0.4 + φ) + 0.60 ,
∗ k + π + φ) + 0.19 ,
(1)
(2)
(3)
where T is the period of the sinusoidal curve measured in the channels, φ is a variable phase of the sinusoidal
curve and k is the channel number. The amplitudes of the three approximation curves and their constant phase shifts
were defined by a method of successive approximations. A necessary condition was that the cross points of the three
sinusoids should correspond to the cross points of the observed curves. It was assumed that the approximation curve
k=1[f (k) − y(k)]2, is minimum. Here f (k) is the
observed curve. A formal least-squares fitting procedure was not applied because it gave unreasonable estimations of
the parameters in view of the large scattering of the experimental data and short segments of the observed curves.
y(k) gives a good fit when the goodness-of-fit parameter, S = P128
We applied the approximation procedure to all the cases presented in the left panel of Fig. 4, varying the frequency
scale T and location φ of the 2.56-MHz band. Initial amplitudes and phases of the curves remained the same. The
results are presented in the right panel of Fig. 4 in the same scale. Comparison of the spectra plotted in both panels
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
5
indicates that the suggested model approximates the observed cases very well. The profile changes at the longitudes of
the three components show appreciable sinusoidal dependence on frequency. The frequency scale of the shape changes
varies from 140 to 450 channels (or from 3 to 9 MHz) for different observations at 112 MHz.
A more detailed picture of the amplitude changes of the average profile with frequency is shown in Fig. 5. The
approximation curves were calculated for T =256 and φ=0. A sinusoidal modulation covers all the three components
of the average profile, but at the longitudes of each component it happens with different phase lags. The lags do not
depend on time. This produces the profile shapes varying with frequency in a similar way for different observations.
Fig. 5 also shows that all the observed profile shapes can be explained by a different location of the receiver bandpass
within a cycle of the sinusoidal curve.
6. Numerical polarization model of the average profile
The frequency picture shown in Fig. 5 is very similar to a Faraday rotation pattern. To clarify whether the observed
effect is due to Faraday rotation, we determined a numerical polarization model of the profile and computed the effects
due to Faraday rotation using the technique that is applied for measurements of linear polarization characteristics of
pulsars at PRAO. The frequency-time dependence of a Faraday rotation pattern was computed using the equation
Ak(i) = a(i) sin (cid:20) 2π
T
∗ k + 2φ(i)(cid:21) + b(i) ,
(4)
where k is the channel number, i is the pulse longitude, T is a period of a sinusoidal curve measured in channels,
φ(i) is a position angle profile, and a(i) is an amplitude of sinusoidal modulation of the polarized component. The
computed values of Ak(i) give the Faraday rotation spectrum of the average profile and corresponding pulse shapes
in each channel. The results are presented in Fig. 6.
In computing Ak(i) we used the published polarization profiles of PSR B0950+08 at low frequencies at 150 and
151 MHz given in Schwarz & Morris (1971) and Lyne et al. (1971). The model profile adopted is presented in Fig. 6a
that shows the average profile in total intensity I(i), the linearly polarized component P (i), and the position angle of
polarization φ(i). The profile for I(i) is normalized to 100 at maximum. In this model, the unresolved component ahead
of the main pulse is 100% linearly polarized (longitude interval from 1◦ to 29◦). The degree of linear polarization, PL,
for the first and second components of the main pulse is 75% and 80%, respectively. Their corresponding longitude
intervals are 29◦ -- 44◦ and 44◦ -- 66◦.
The amplitude of sinusoidal curve a(i) in Eq. (4) was defined by the relation a(i) = P (i)/2, where P (i) = I(i)∗PL(i)
is the intensity of the polarized component at longitude i. The mean level of the pulse intensity b(i) at longitude i was
found from b(i) = I(i)/2. The period of the sinusoidal curve was assumed equal to T =256 channels, the same as in
Fig. 5. According to Lyne et al. (1971), the position angle φ swings 160◦ through the profile. This gives the phase offset
between the Faraday spectra of two components of the main pulse of about -0.4 radian, i.e. the same as in Eq. (2). To
obtain the phase offset between the spectra of the first and third components equal to ≈ π, as specified in Eq. (3) and
Fig. 5, we should suppose that the position angles for these components are orthogonal. The finally adopted behavior
of the position angle is given in Fig. 6a in the top panel. Here the position angle swings 200◦ through the profile and
decreases across the profile with a higher rate of decrease in the range of the third component. This gives about 90◦
displacement from the position angle of the first component.
The resulting frequency dependence of the pulse profile at the longitudes of the pulse peaks of the three components
is given in Fig. 6b. Comparison with Fig. 5 shows that the computed Faraday rotation spectra are generally in good
agreement with the observed frequency dependence of the profile. The values of the phase offsets between the three
sinusoidal curves are identical in both plots. The values of the amplitudes are slightly different. The amplitudes of the
observed sinusoidal curves for the two components of the main pulse are about half that shown in Fig. 6b. Possibly, the
procedure of removing the effects of scintillations, which was mentioned in §5, could partially decrease the amplitude
of the wave, unrelated to scintillations. The amplitude of the third component ahead of the main pulse is 4 times
greater than that plotted in Fig. 6b. Possibly, at 112 MHz this component is stronger than it was assumed in our
model. We do not know the true profile in total intensity at 112 MHz and therefore the assumed ratios of amplitudes
in the various profile components are approximate.
The corresponding profile shapes computed for individual channels are given in Fig. 6c. Comparison with Fig. 1
shows that the computed profiles are in good agreement with the observed ones. We clearly see the noticeable feature
which characterizes the observed modulation of the pulse profiles. This is a transition state between the first and second
dominant component of the main pulse. This state has a lower S/N ratio and shows the presence of the well-resolved
third component ahead of the main pulse. The well-resolved third component does not exist. The transformation of the
unresolved component of the pulse profile to the well-resolved one is a result of polarization effects. The simulation of
the transition state was successful when the degree of linear polarization taken was very high, more than 80%. Hence,
we conclude that the observed profile changes are a result of the Faraday rotation effect.
6
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
7. Narrowband changes of the average profile at lower frequencies of 88, 62 and 41 MHz
Low-frequency observations of the pulsar were carried out to examine sinusoidal modulation of the pulse profile in the
wide frequency range and to estimate the polarization effects corresponding to our observations. As mentioned before,
the observed effects at 112 MHz are about 3 times stronger than those predicted by the interstellar Faraday rotation.
Fig. 7 displays the average pulse profile at 88.57 MHz. The profile exhibits a well-resolved double shape with a
separation between the components of about 11◦. As we can see from this plot, the profile shape varies markedly
across the bandpass and these changes are similar to the changes observed at the higher frequency of 112 MHz. The
transition state between the first and second dominant components was not recorded during this observation, but one
can guess its presence. The frequency interval of the shape changes can be estimated to be about 3 MHz.
A similar picture of the narrowband changes of the pulse profile is observed at 62.15 MHz as is shown in Fig. 8.
Here the profile is obtained by addition of 2200 pulse periods within eight 320-kHz bands. The average profile has two
well-resolved components with a separation of 15◦. Obviously, the profile shape varies across the bandpass and the
frequency structure of the profile is similar to that observed at higher frequencies of 88 and 112 MHz. A noticeable
feature of the shape changes is the presence of two transition states between the first and second dominant components,
which are clearly visible in the plot. The frequency interval of the profile changes of approximately 1.2 MHz.
Average profiles of the pulsar at 41.07 MHz do not show a frequency structure similar to that observed at higher
frequencies. The profiles have a lower S/N ratio and much lower time resolution of 3 ms, caused by the 7-ms pulse
broadening due to the dispersion effect in one 20-kHz band. The profiles would be expected to be modulated by the
interstellar Faraday rotation with the frequency scale of the order of 900 kHz. To find periodicities in the amplitude
variation of the pulse profile over the bandpass of 2.56 MHz, the power spectrum was computed with the use of Fourier
transform. For this procedure, we used strong individual pulses. It is well known that individual pulses are often more
highly polarized than average profiles. The results are given in Fig. 9 for three different epochs. For each observation
two realizations of the power spectrum are presented.
There is a distinct spectral feature at around 0.005 kHz−1, which is clearly visible in the spectra for the three
observations. This feature corresponds to the periodicity of approximately 200 kHz. To state with certainty the
presence of this spectral feature we averaged the spectra together and presented the resulting spectrum in the lower
panel of Fig. 9. As is seen, the spectral feature at 0.005 kHz−1 dominates over the noisy fluctuations. This is a strong
indication of the presence of sinusoidal modulation of the pulse profile at frequency of 41 MHz, too.
In this section we analyzed the narrowband profile changes in the low-frequency range from 41 to 88 MHz and
concluded that the shape changes are similar to those observed at 112 MHz. The scale of the observed profile changes
is frequency dependent and varies from 3 -- 9 MHz at 112 MHz to 0.2 MHz at 41 MHz.
8. Timing residuals for PSR B0950+08
The pulsar catalog (Taylor et al. 1993) quotes the period and the period derivative obtained by Gullahorn & Rankin
(1978) at the epoch of 1972.5. Errors in these parameters alone lead to ambiguities in the counts of the pulse number for
the 1996 -- 2001 observing session. Our measurements of the average profiles were used to improve the timing parameters
for PSR B0950+08.
We analyzed the arrival times for only those average profiles whose shape looks double. They were chosen out of
the full data set obtained at 102.5 and 112 MHz between 1996 and 2001. All good-quality double profiles were summed
to produce an average profile with a high S/N ratio which was used as a template. The topocentric arrival times were
calculated by cross-correlating the observed profiles with this template. Timing parameters were determined using the
TEMPO software package1 and the JPL DE200 ephemeris. The topocentric arrival times were referred to the solar
system barycenter using a position and a proper motion given by Fomalont et al. (1992) and Brisken et al. (2000),
respectively. A second-order polynomial describing spin-down pulsar behavior was fitted to the barycentric arrival
times to obtain residuals from a timing model. Residuals were used to compute differential corrections to the initial
parameter values.
The best-fitting values of the period and the period derivative are presented in Table 2. A large scatter of timing
data relative to the spin-down pulsar model has not allowed us to improve the astrometric parameters. Fig. 10 displays
the timing residuals from the pulsar relative to their best-fit spin-down model during the period 1996 -- 2001. The rms
timing residual after the fit was of the order of 1300 µs with a timing error of about 90 µs. At higher frequencies,
above 400 MHz, where the pulsar exhibits a single pulse profile, the rms timing residual of PSR B0950+08 is much less
and is estimated at 200 µs with a timing uncertainty of the order of 90 -- 200 µs in a 4 -- 7 year span of data (Gullahorn
& Rankin 1978; Helfand et al. 1980). Apparently, the increase of the rms residual observed at 112 MHz is due to the
profile shape changes which cause fluctuations of the pulse arrival times and make an additional noise contribution to
the timing residual.
1 http://www.atnf.csiro.au/research/pulsar/timing/tempo
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
7
Table 2. Measured timing parameters for PSR B0950+08
Parameter
Valuea
Period, P (s)
Period derivative,
Epoch of period (MJD)
P (10−15 s s−1)
0.253065240711(4)
0.23028(5)
50190.7360
a The errors are given in units of the last quoted digit
9. Discussion
We have found that the average pulse profile for PSR B0950+08 observed with a single linear polarization is frequency
variable all over the range 41 -- 112 MHz. The narrowband structure of the pulse shape is associated with sinusoidal
modulation of the pulsar emission. This process is frequency dependent. According to our measurements, the interval
of modulation at different observing frequencies is about 3 -- 9 MHz at 111.87 MHz, 3 -- 4 MHz at 88.57 MHz, 1 -- 1.4 MHz
at 62.15 MHz and 0.19 -- 0.21 MHz at 41.07 MHz. Hence, the dependence of the modulation interval on frequency is
well described by a power low with index close to 3 in the range 41 -- 112 MHz. A similar frequency structure should
be observed in the presence of the Faraday rotation effect. The presence of Faraday rotation is also confirmed by a
numerical simulation of the polarization effects for the pulse profiles at 112 MHz.
The frequency interval of Faraday rotation of the plane of polarization of linear polarized emission ∆Fπ at the
observing frequency ν is determined by rotation in the position angle θ through π and is calculated from
∆Fπ =
π ν 3
2 RM c2 = 17.48 ν 3 / RM ,
(5)
where ∆Fπ is in MHz, ν is in hundreds of MHz, the rotation measure RM is in rad m−2 and c is the light velocity.
Using relation (5), we can estimate the rotation measure RM corresponding to our measurements. For the mean
values of ∆Fπ equal to 6, 3, 1.2 and 0.21 MHz at the frequencies of 111.87, 88.57, 62.15 and 41.07 MHz respectively, the
rotation measure RM will be approximately 4 rad m−2. Taking into account variability of ∆Fπ from one observation
to another, we derive the values of RM in the range 3 -- 6 rad m−2. The derived values of RM are very large compared
to the rotation measure RM = 1.35 rad m−2 measured for PSR B0950+08 at frequencies above 400 MHz (Hamilton
& Lyne 1987; Taylor et al. 1993).
Thus, we have shown that the observed narrowband profile changes of the pulsar B0950+08 are a result of the
Faraday rotation effect and could be explained as an artifact of observations of linearly polarized pulsar emission
with a linearly polarized antenna. We have obtained a new value of the rotation measure corresponding to this effect,
RM ≈ 4 rad m−2. This leads us to conclude that the published value of RM is incorrect and is actually 3 times
greater.
Acknowledgements. We wish to thank the staff of PRAO for assistance in the observations. This work was supported by the
Russian Foundation For Basic Research (project No. 00-02-17447). The authors are grateful to the referee for numerous helpful
comments and suggestions.
References
Brisken, W. F., Benson, J. M., Beasley, A. J., et al. 2000, ApJ, 541, 959
Fomalont, E. B., Goss, W. M., Lyne, A. G., Manchester, R. N., & Justtanont, K. 1992, MNRAS, 258, 497
Gullahorn, G. E., & Rankin, J. M. 1978, AJ, 83, 1219
Hamilton, P. A., & Lyne, A. G. 1987, MNRAS, 224, 1073
Hankins, T. H., & Cordes, J. M. 1981, ApJ, 249, 241
Hankins, T. H., Izvekova, V. A., Malofeev, V. M., et al. 1991, ApJ, 373, L17
Helfand, D. J., Manchester, R. N., & Taylor, J. H. 1975, ApJ, 198, 661
Helfand, D. J., Taylor, J. H., Backus, P. R., & Cordes, J. M. 1980, ApJ, 237, 206
Kuzmin, A. D., Izvekova, V. A., Shitov, Yu. P., et al. 1998, A&AS, 127, 355
Lyne, A. G., & Rickett B. J. 1968, Nature, 218, 326
Lyne, A. G., Smith, F. G., & Graham D. A. 1971, MNRAS, 153, 337
Rathnasree, N., & Rankin, J. M. 1995, ApJ, 452, 814
Schwarz, U. J., & Morris D. 1971, Astrophys. Lett., 7, 185
Shitov, Yu. P. 1971, AZh, 48, 638
Shitov, Yu. P. 1972, AZh, 49, 470
8
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
Smirnova, T. V., & Shabanova, T. V. 1992, Soviet Astron, 36, 628
Suleymanova, S. A. 1989, Linear polarization of the average pulses of pulsars at 102.5, 60 and 40 MHz. In Pulsars, ed. A. D.
Kuzmin, Proc. Lebedev Physical Institute, vol. 196, 55
Taylor, J. H., Manchester, R. N., & Lyne, A. G. 1993, ApJS, 88, 529
Udaltsov, V. A., & Zlobin, V. N. 1974, A&A, 37, 21
Vitkevich, V. V., & Shitov, Yu. P. 1970, Nature, 226, 1235
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
9
'#&
B ##0
B$0
J "&I
0ECD
.HAGKA?O
&''%
"'''
%"''%
!''$
$"''%
!!''&
%!''&
& '''
&''%
#
=
>
?
@
A
B
C
D
M
.HAGKA?O
=
#
#
CEJK@A@ACHAAI
Fig. 1. A sequence of shapes of the average profile continuously varying between the main states of the double profile.
The profiles were chosen from the full data set obtained between 1996 and 2001 at frequencies around 102 and 112
MHz. Shapes from (b) to (g) show a triple profile with a well-resolved component ahead of the main pulse located
at longitudes from 150◦ to 180◦. The sample interval is 0.◦3 of longitude. All the profiles are aligned in time and
normalized to unity.
10
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
'#&
B$"0
%'
=
&%
!
#'
''#
&%
!
#'
''#
&%
!
#'
''#
'
>
$
?
#
#
CEJK@A@ACHAAI
Fig. 2. The changes of the average profile with frequency for three different epochs. A set of 12 plotted profiles covers
the full cycle of the shape changes. Panels (a) and (c) demonstrate the transition state between the first and second
dominant components of the main pulse (marked by arrow). The sample interval is 0.◦7 of longitude. All the profiles
are normalized to unity.
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
11
%'
B$0
'#&
=
'
B$0
'#&
>
&%
%
##
!'
!
%
'
%#
#'
"!
%
''#
'%'
'$!
'"%
&%
%
##
!'
!
%
'
%#
#'
"!
%
''#
'%'
'$!
'"%
#
#
#
#
CEJK@A@ACHAAI
CEJK@A@ACHAAI
Fig. 3. The frequency changes of the average profile between 160-kHz bands for two different epochs. To watch the
profile frequency dependence inside one cycle of changes, the right hand sequence should be considered as an extension
of the left hand one. The transition state between the first and second dominant components is marked by the arrow.
All the profiles are normalized to unity.
12
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
'
=N1
'#&
'
6!?D=
=N11
"
=N1
'
=N1
$
I
J
E
K
O
H
=
H
J
E
>
H
=
O
J
E
I
A
J
1
=N11
=N11
=N11
=N1
=N11
%'
=N1
"
6"#?D=
'
6 ?D=
$
6&?D=
%'
6"?D=
!
'$
+D=AK>AH
$"
&
!
'$
+D=AK>AH
$"
&
Fig. 4. The frequency structure of the average profile at the longitudes of the pulse peaks of the three components
for different epochs: the observed spectra (Left), the model spectra (Right). The spectrum for the first component
of the main pulse is marked by filled circles and the word "maxI", the spectrum for the second component is marked
by open circles and the word "maxII", the spectrum for the third component is marked by crosses.
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
13
6 #$
=N11
=N1
O
J
E
I
A
J
1
&
#$
!&"
#
+D=AK>AH
Fig. 5. The model picture of the amplitude changes of the average profile with frequency at the longitudes of the
pulse peaks of the three components for T = 256 and φ=0. The spectrum is marked by filled circles and the word
"maxI" for the first component, open circles and the word "maxII" for the second component and crosses for the third
one.
14
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
?
'#&
=
!
"
,ACHAAI
#
$
6 #$
>
O
?
A
K
G
A
H
.
C
A
@
)
2
&
'
O
J
E
I
A
J
1
O
J
E
I
A
J
1
&
!&"
+D=AK>AH
#$
#
"
,ACHAAI
$
Fig. 6. A numerical study of polarization effects. a) The model profile adopted shows average profile in total intensity
(solid), the linearly polarized component (dotted), and the position angle of polarization (upper panel). b) The
computed Faraday rotation spectra at the longitudes of the pulse peaks of the three components: 14◦ (crosses), 38◦
(filled circles) and 49◦ (open circles). c) The computed profile shapes for individual channels. All the profiles are
normalized to unity. A transition state between the first and second dominant components of the main pulse is marked
by an arrow.
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
15
&
B&&#%0
B$"0
'#&
&&#%
&%'!
&% '
&$$#
#
#
CEJK@A@ACHAAI
Fig. 7. The pulse profile changes at the frequency of 88.57 MHz. The upper profile is a sum of 3700 pulsar periods over
a total 2.56-MHz band. The lower profiles are that over four 640-kHz bands. The sample interval is 1.◦2 of longitude.
All the profiles are normalized to unity.
16
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
'
B$ #0
B! 0
'#&
$ #
$'
#''
#
#
CEJK@A@ACHAAI
Fig. 8. The pulse profile changes at the frequency of 62.15 MHz. The upper profile is a sum of 2200 pulsar periods
over a total 2.56-MHz band. The lower profiles are that over eight 320-kHz bands. A transition state between the first
and second dominant components is marked by an arrow. The sample interval is 1.◦2 of longitude. All the profiles are
normalized to unity.
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
17
$"
%"
&"
=LAH=CA
H
A
M
2
=
H
J
?
A
F
5
" & $
.HAGKA?O0
Fig. 9. The power spectra of the amplitude variations of individual pulses for PSR B0950+08 at 41 MHz for three
different observations. The spectral feature at around 0.005 kHz−1 is well visible in the spectra for all the three
observations. In the average spectrum shown in the lower panel this feature dominates over the noisy fluctuations.
18
Shabanova and Shitov: Linearly polarized radiation from PSR B0950+08
'#&
I
=
K
@
E
I
A
4
"
"
#
#
,
#
Fig. 10. Post-fit timing residuals for PSR B0950+08 between 1996 April and 2001 July.
|
0802.3823 | 1 | 0802 | 2008-02-26T16:03:07 | Space VLBI polarimetry of IDV sources: Lessons from VSOP and prospects for VSOP2 | [
"astro-ph"
] | To locate and image the compact emission regions in quasars, which are closely connected to the phenomenon of IntraDay Variability (IDV), space VLBI observations are of prime importance. Here we report on VSOP observations of two prominent IDV sources, the BL Lac objects S5 0716+714. To monitor their short term variability, these sources were observed with VSOP at 5 GHz in several polarisation sensitive experiments, separated in time by one day to six days, in autumn 2000. Contemporaneous flux density measurements with the Effelsberg 100m radio telescope were used to directly compare the single dish IDV with changes of the VLBI images. A clear IDV behaviour in total intensity and linear polarization was observed in 0716+714. Analysis of the VLBI data shows that the variations are located inside the VLBI core component of 0716+714. In good agreement with the single-dish measurements, the VLBI ground array images and the VSOP images, both show a decrease in the total flux density of ~20 mJy and a drop of ~5 mJy in the linear polarization of the VLBI core. No variability was found in the jet. From the variability timescales we estimate a source size of a few micro-arcseconds and brightness temperatures exceeding 10^15 K. Independent of whether the interpretation of the IDV seen in the VLBI core is source intrinsic or extrinsic a lower limit of T_B > 2x10^12 K is obtained by model fitting of the VLBI-core. Our results show that future VSOP2 observations should be accompanied by a single dish monitoring not only to discriminate between source-extrinsic and source-intrinsic effects but to allow also a proper calibration and interpretation of ultra-high resolution VSOP2 images. | astro-ph | astro-ph |
**FULL TITLE**
ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION**
**NAMES OF EDITORS**
Space VLBI polarimetry of IDV sources: Lessons from
VSOP and prospects for VSOP2
U. Bach1, T.P. Krichbaum1, S. Bernhart1
A. Kraus1, L. Fuhrmann1, A. Witzel1 & J.A. Zensus1
,
2, C.M.V. Impellizzeri1,
1Max-Planck-Institut fuer Radioastronomie, Bonn, Germany
2IGG, University of Bonn, Germany
Abstract.
To locate and image the compact emission regions in quasars, which are
closely connected to the phenomenon of IntraDay Variability (IDV), space VLBI
observations are of prime importance. Here we report on VSOP observations
of two prominent IDV sources, the BL Lac objects S5 0716+714. To monitor
their short term variability, these sources were observed with VSOP at 5 GHz
in several polarisation sensitive experiments, separated in time by one day to
six days, in autumn 2000. Contemporaneous flux density measurements with
the Effelsberg 100 m radio telescope were used to directly compare the single
dish IDV with changes of the VLBI images. A clear IDV behaviour in total
intensity and linear polarization was observed in 0716+714. Analysis of the
VLBI data shows that the variations are located inside the VLBI core component
of 0716+714. In good agreement with the single-dish measurements, the VLBI
ground array images and the VSOP images, both show a decrease in the total flux
density of ∼ 20 mJy and a drop of ∼ 5 mJy in the linear polarization of the VLBI
core. No variability was found in the jet. These findings are supported by VLBA
observations of five IDV sources, including 0716+714, in December 2000, that
show a similar behaviour. From the variability timescales we estimate a source
size of a few micro-arcseconds and brightness temperatures exceeding 1015 K.
Independent of whether the interpretation of the IDV seen in the VLBI core is
source intrinsic or extrinsic a lower limit of TB > 2 × 1012 K is obtained by model
fitting of the VLBI-core. Our results show that future VSOP2 observations
should be accompanied by a single dish monitoring not only to discriminate
between source-extrinsic (interstellar scintillation) and source-intrinsic effects
but to allow also a proper calibration and interpretation of ultra-high resolution
VSOP2 images.
1.
Introduction
Since the discovery of intraday variability (IDV, i.e. flux density and polarization
variations on time scales of less than 2 days) about 20 years ago (Witzel et al.
1986; Heeschen et al. 1987), it has been shown that IDV is a common phe-
nomenon among extra-galactic compact flat-spectrum radio sources. It is de-
tected in a large fraction of this class of objects (e.g. Quirrenbach et al. 1992;
Kedziora-Chudczer et al. 2001; Lovell et al. 2003, see also D. Jauncey et al.;
H. Bignall et al., these proceedings). The occurrence of IDV appears to be
correlated with the compactness of the VLBI source structure on milliarcsec-
ond scales: IDV is more common and more pronounced in objects dominated
1
2
by a compact VLBI core than in sources that show a prominent VLBI jet. In
parallel to the variability of the total flux density, variations in the linearly polar-
ized flux density and the polarization angle have been observed in many sources
(e.g. Quirrenbach et al. 1989; Kraus et al. 1999, 2003; Qian & Zhang 2004). The
common explanation for the IDV phenomenon at cm-wavelength is nowadays in-
terstellar scattering (e.g. Rickett et al. 1995; Rickett 2001b). On the other hand
some effects remain that cannot be easily explained by interstellar scintillation
and that are probably caused by relativistic jet physics (e.g. Qian et al. 1996,
2002; Qian & Zhang 2004). For example the correlated intra-day variability be-
tween radio and optical wavelengths, which is observed in sources like 0716+714
and 0954+658 (e.g. Wagner et al. 1990; Quirrenbach et al. 1991; Wagner et al.
1996) and the recent detection of IDV at millimetre wavelengths in 0716+714
(Krichbaum et al. 2002; Kraus et al. 2003; Agudo et al. 2006; Fuhrmann et al.
2008) suggests that at least part of the observed IDV has a source-intrinsic
origin.
Independent of the physical cause of IDV, it is obvious that IDV sources
must contain one or more ultra-compact emission regions. Using scintillation
models, typical source sizes of a few ten micro-arcseconds have been derived (e.g.
Rickett et al. 1995; Dennett-Thorpe & de Bruyn 2002; Bignall et al. 2003). In
the case of source intrinsic variability and when using the light-travel-time argu-
ment, even smaller source sizes of a few micro-arcseconds are obtained. In this
case it implies apparent brightness temperatures of up to 1018−19 K (in excep-
tional cases up to 1021 K), far in excess of the inverse Compton limit of 1012 K
(Kellermann & Pauliny-Toth 1969). These high apparent brightness tempera-
tures can be reduced e.g. by relativistic beaming with high Doppler-factors (e.g.
Qian et al. 1991, 1996; Kellermann 2002).
The motivation of this VLBI monitoring, therefore, was to find how rapid
structural variability on sub-mas-scales occurs and where the IDV component
is located in the jet. An array of 12 antennas consisting of the 10 stations
of NRAO's VLBA, the 100 m radio telescope of the Max-Planck-Institut fur
Radioastronomie in Effelsberg (Germany), and the 8 m HALCA antenna of the
VSOP was used to follow the short-term variability of 0716+714 and 0954+658
at 5 GHz in autumn 2000. During short gaps in the VLBI schedule the Effelsberg
antenna was used to measure the light curve of a calibrator and our target
sources. A detailed description of the experiment and data reduction can be
found in Bach et al. (2006). Here we will concentrate on the results of 0716+714,
the results of 0954+658 will be presented by Bernhart et al. (in prep.).
2. Results and Discussion
The ground array and the VSOP images of 0716+714 show a bright core and a
jet oriented to the north (Fig. 1). The linear polarization images indicate that
the jet magnetic field is perpendicular to the jet axis. Compared to the jet axis,
the electric vector position angle in the core is misaligned by around 60 ◦. This
is explained either by opacity effects in the core region or by a curved jet. Here,
jet curvature is supported by recent high resolution 3 mm VLBI observations
that show the inner jet structure (r < 0.1 mas) at a similar position angle as the
EVPA in the core at 6 cm wavelength (Bach et al. 2006).
3
Oct. 5
1.2
>
P
<
P
/
1
0.8
0836+710
0716+714
)
s
a
m
(
.
l
c
e
D
e
v
i
t
a
l
e
R
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
-0.5
-1.0
Oct. 4
)
s
a
m
(
.
l
c
e
D
e
v
i
t
a
l
e
R
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
-0.5
-1.0
11821.5
11822
11822.5
JD-2440000
11823
2.5
2.0
1.5
1.0
0.5
0.0 -0.5 -1.0 -1.5 -2.0
Relative R.A. (mas)
2.5
2.0
1.5
1.0
0.5
0.0 -0.5 -1.0 -1.5 -2.0
Relative R.A. (mas)
Figure 1.
Left: Single dish polarization variability of 0716+41 on Oct. 4
and 5. Middle and right: VSOP contour maps of Stokes I of 0716+714
with polarization vectors superimposed (1 mas corresponds to 6.7 mJy/beam).
Contours start at 1.8 mJy/beam increasing in steps of 2. A clear drop in
polarized flux (shorter vectors) and a rotation by 10◦ is visible in the core
region from Oct. 4 to 5.
Simultaneous flux-density measurements with the 100 m Effelsberg tele-
scope during the VSOP observations revealed variability in total intensity (∼
5 %) and in linear polarization (up to ∼ 40 %) accompanied by a rotation of the
polarization angle by up to 15 ◦. The analysis of the VLBI data shows that the
intra-day variability in 0716+714 is associated with the VLBI-core region and
not with the milli-arcsecond jet. Both the ground array and the VSOP maps
show a similar decrease of the flux densities in total intensity and linear polar-
ization of the core component, which is in good agreement with the variations
seen with the Effelsberg radio-telescope (more details are given in Bach et al.
2006). These findings could be confirmed by VLBA observations of five IDV
sources, including 0716+714, a few month later in December 2000, that show
a similar behaviour (Impellizzeri et al. in perp.). In this, 0716+714 displays a
behaviour that is similar to what was previously observed in the IDV sources
0917+624 and 0954+658, where components in or near the VLBI core region
were also made responsible for the IDV (Gabuzda et al. 2000b). Over the time
interval of our VSOP observations, no rapid variability in the jet was observed
and we cannot confirm the variability outside the core and in the jet found by
Gabuzda et al. (2000a).
The simultaneous variation of the polarization angle with the polarized
intensity in the core suggests that the variations might be the result of the
sum of the polarization of more than one compact sub-component on scales
smaller than the beam size (multi-comp. model). Assuming a redshift of z = 0.3
(Wagner et al. 1996) and that these variations are intrinsic to the source, we
derived brightness temperatures of ∼ 3 × 1015 K to ∼ 1016 K. Doppler factors of
> 20 are needed to bring these values down to the inverse-Compton limit. These
numbers agree with the observed speeds in the jet if the angle to the line of sight
is very small (θ < 2◦), as already proposed by Bach et al. (2005). Because of
the unknown redshift, the derived speeds and brightness temperatures represent
only lower limits.
4
However, interstellar scintillation effects could also explain the IDV seen in
the VLBI core, if the core region consists of several compact and polarized sub-
components, with sizes of a few ten micro-arcseconds. To explain the observed
polarization variations, the sub-components must scintillate independently in a
different manner, which means that they must have slightly different intrinsic
sizes and intrinsic polarization (e.g. Rickett 2001b,a).
Independent of whether the interpretation of the IDV seen in the VLBI core
is source intrinsic or extrinsic, the space-VLBI limit to the core size (< 0.1 mas)
gives a robust lower limit to the brightness temperature of ≥ 2 × 1012 K and
therewith exceeds the inverse-Compton limit. This implies a lower limit to the
Doppler factor of about ≥ 4 and, independent of the model we use to explain
the variability, relativistic beaming must play a role.
The increased sensitivity, higher resolution, and frequency flexibility of
VSOP-2 will provide a powerful tool to look even deeper into the core region
and to distinguish which fraction of the IDV is due to source intrinsic variations
and which is caused by the interstellar medium. Variability surveys like MASIV
(Lovell et al. 2003) revealed that IDV is present in a huge number of sources and
therefore it seems advisable that future VSOP2 experiments are accompanied
by a simultaneous flux monitoring to allow a proper calibration of the data and
to interpret possible structural changes on short time scales.
References
Agudo I., Krichbaum T. P., Ungerechts H., et al., 2006, A&A 456, 117
Bach U., Krichbaum T. P., Kraus A., Witzel A., Zensus J. A., 2006, A&A 452, 83
Bach U., Krichbaum T. P., Ros E., et al., 2005, A&A 433, 815
Bignall H. E., Jauncey D. L., Lovell J. E. J., et al., 2003, ApJ 585, 653
Dennett-Thorpe J., de Bruyn A. G., 2002, Nat 415, 57
Fuhrmann L., et al., 2008, A&A in press.
Gabuzda D. C., et al., 2000a, MNRAS 313, 627
Gabuzda D. C., et al., 2000b, MNRAS 315, 229
Heeschen D. S., Krichbaum T. P., Schalinski C. J., Witzel A., 1987, AJ 94, 1493
Kedziora-Chudczer L. L., et al., 2001, MNRAS 325, 1411
Kellermann K. I., 2002, Publications of the Astronomical Society of Australia 19, 77
Kellermann K. I., Pauliny-Toth I. I. K., 1969, ApJ 155, L71
Kraus A., Krichbaum T. P., Wegner R., et al., 2003, A&A 401, 161
Kraus A., Witzel A., Krichbaum T. P., 1999, New Astronomy Review 43, 685
Krichbaum T. P., Kraus A., Fuhrmann L., Cim`o G., Witzel A., 2002, PASA 19, 14
Lovell J. E. J., Jauncey D. L., Bignall H. E., et al., 2003, AJ 126, 1699
Qian S., Kraus A., Zhang X., et al., 2002, ChJA&A 2, 325
Qian S., Li X., Wegner R., Witzel A., Krichbaum T. P., 1996, ChJA&A 20, 15
Qian S., Zhang X., 2004, ChJA&A 4, 37
Qian S. J., Quirrenbach A., Witzel A., et al., 1991, A&A 241, 15
Quirrenbach A., Witzel A., Kirchbaum T. P., et al., 1992, A&A 258, 279
Quirrenbach A., Witzel A., Qian S. J., et al., 1989, A&A 226, L1
Quirrenbach A., Witzel A., Wagner S., et al., 1991, ApJ 372, L71
Rickett B. J., 2001a, Ap&SS 278, 129
Rickett B. J., 2001b, Ap&SS 278, 5
Rickett B. J., Quirrenbach A., Wegner R., et al., 1995, A&A 293, 479
Wagner S., Sanchez-Pons F., Quirrenbach A., Witzel A., 1990, A&A 235, L1
Wagner S. J., Witzel A., Heidt J., et al., 1996, AJ 111, 2187
Witzel A., et al., 1986, Mit. d. Astro. Ges. Hamburg 65, 239
|
astro-ph/0503110 | 1 | 0503 | 2005-03-04T20:09:07 | The effective temperature scale of FGK stars. II. Teff : color : [Fe/H] calibrations | [
"astro-ph"
] | We present up-to-date metallicity-dependent temperature vs. color calibrations for main sequence and giant stars based on temperatures derived with the infrared flux method (IRFM). Seventeen colors in the following photometric systems: UBV, uvby, Vilnius, Geneva, RI(Cousins), DDO, Hipparcos-Tycho, and 2MASS, have been calibrated. The spectral types covered range from F0 to K5 (7000 K<Teff<4000 K) with some relations extending below 4000 K or up to 8000 K. Most of the calibrations are valid in the metallicity range -3.5<[Fe/H]<0.4, although some of them extend to as low as [Fe/H]=-4.0. All fits to the data have been performed with more than 100 stars; standard deviations range from 30 K to 120 K. Fits were carefully performed and corrected to eliminate the small systematic errors introduced by the calibration formulae. Tables of colors as a function of Teff and [Fe/H] are provided. (Abridged) | astro-ph | astro-ph | To Appear in ApJ
Preprint typeset using LATEX style emulateapj v. 6/22/04
5
0
0
2
r
a
M
4
1
v
0
1
1
3
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
THE EFFECTIVE TEMPERATURE SCALE OF FGK STARS. II. Teff : COLOR : [Fe/H] CALIBRATIONS
Department of Astronomy, University of Texas at Austin, RLM 15.306, TX 78712-1083
Iv´an Ram´ırez1
Department of Astronomy, California Institute of Technology, MC 105 -- 24, Pasadena, CA 91125
and
Jorge Mel´endez1
Preprint version October 30, 2018
ABSTRACT
We present up-to-date metallicity-dependent temperature vs. color calibrations for main sequence
and giant stars based on temperatures derived with the infrared flux method (IRFM). Seventeen colors
in the following photometric systems: U BV , uvby, Vilnius, Geneva, RI(Cousins), DDO, Hipparcos-
Tycho, and 2MASS, have been calibrated. The spectral types covered range from F0 to K5 (7000 K&
Teff &4000 K) with some relations extending below 4000 K or up to 8000 K. Most of the calibrations
are valid in the metallicity range −3.5 & [Fe/H] & 0.4, although some of them extend to as low
as [Fe/H] ∼ −4.0. All fits to the data have been performed with more than 100 stars; standard
deviations range from 30 K to 120 K. Fits were carefully performed and corrected to eliminate the
small systematic errors introduced by the calibration formulae. Tables of colors as a function of Teff
and [Fe/H] are provided. This work is largely based on the study by A. Alonso and collaborators and
thus our relations do not significantly differ from theirs except for the very metal-poor hot stars. From
the calibrations, the temperatures of 44 dwarf and giant stars with direct temperatures available are
obtained. The comparison with direct temperatures confirms our finding in Part I that the zero point
of the IRFM temperature scale is in agreement, to the 10 K level, with the absolute temperature scale
(that based on stellar angular diameters) within the ranges of atmospheric parameters covered by
those 44 stars. The colors of the Sun are derived from the present IRFM Teff scale and they compare
well with those of five solar analogs. It is shown that if the IRFM Teff scale accurately reproduces
the temperatures of very metal-poor stars, systematic errors of the order of 200 K, introduced by
the assumption of (V − K) being completely metallicity-independent when studying very metal-
poor dwarf stars, are no longer acceptable. Comparisons with other Teff scales, both empirical and
theoretical, are also shown to be in reasonable agreement with our results, although it seems that
both Kurucz and MARCS synthetic colors fail to predict the detailed metallicity dependence, given
that for [Fe/H] = −2.0, differences as high as ∼ ±200 K are found.
Subject headings: stars: atmospheres -- stars: fundamental parameters -- stars: abundances
1. INTRODUCTION
Theoretical calculations in astrophysics predict rela-
tions between physical quantities such as effective tem-
perature (Teff), luminosity (L), and stellar radius. These
quantities are, in general, not directly measurable, and so
their correspondence with observational quantities such
as colors and magnitudes play a crucial role in the inter-
pretation of the results. The importance and usefulness
of such correlations are discussed in the following para-
graphs.
Stellar chemical compositions are derived from the
comparison of synthetic and observed spectra, either by
line-profile fitting (e.g. Hill et al. 2002, Allende Prieto
et al. 2004, Sneden et al. 2003) or equivalent width
matches (e.g. Reddy et al. 2003, Takeda et al. 2002).
Both methods require both Teff and log g (the surface
gravity) as input parameters and so their uncertainties
are reflected in the abundances derived. The accurate
determination of effective temperatures is thus a critical
step in any abundance analysis.
1 Affiliated with the Seminario Permanente de Astronom´ia y
Ciencias Espaciales of the Universidad Nacional Mayor de San Mar-
cos, Peru
The theory of stellar evolution deals with the evolution
in time of the fundamental stellar physical parameters,
which is very well illustrated in the Teff vs. L plane
(the theoretical HR diagram). Theoretical isochrones
and evolutionary tracks are the final products of these
calculations (e.g. Girardi et al. 2002, Yi et al. 2003).
The transformation of the effective temperature axis into
a color axis, along with a transformation of the luminos-
ity axis into an absolute magnitude axis (i.e. the trans-
formation of the Teff vs. L plane into a color-magnitude
diagram), allow observations to be compared with theo-
retical results, leading to a better understanding of the
systems studied, or to the test of the models themselves.
There is a continuous feedback between theory and ob-
servation through these kinds of transformations.
A problem of particular interest, whose resolution may
be partly in the adopted Teff scale is that of the primor-
dial lithium abundance, A(Li), which is derived from the
observation of metal-poor dwarfs. Ryan et al.
(1999)
determined lithium abundances in very metal-poor stars
(−3.6 < [Fe/H] < −2.3) employing the temperature cal-
ibration of Magain (1987), whose metallicity dependence
was derived with only one star (HD 140283, [Fe/H] =
−2.5) in the metallicity range used by Ryan et al., and
2
Ram´ırez & Mel´endez
the few other metal-poor stars included in the calibra-
tion have [Fe/H] > −2.15. Ryan et al. claim that the
lithium abundance in halo stars depends on metallicity,
and extrapolating to zero metals they derive a primor-
dial abundance of A(Li) = 2.0 dex. On the other hand,
using the standard theory of Big Bang nucleosynthesis
and the baryon-to-photon ratio determined from WMAP
(Wilkinson Microwave Anisotropy Probe), a primordial
lithium abundance of 2.6 dex is derived (Romano et al.
2003, Coc et al. 2004), much higher than the lithium
abundance obtained by Ryan et al. (1999).
Bonifacio & Molaro (1997) found that when the Teff
scale from the infrared flux method (IRFM) is employed,
the lithium abundance in very metal-poor stars does not
depend on metallicity, and a primordial A(Li) = 2.24
dex is obtained, significantly higher than the primordial
abundance proposed by Ryan et al., but still lower than
the abundance suggested by the WMAP data. Since the
temperature is the key parameter to obtain lithium abun-
dances, we have included very metal-poor F and G dwarfs
in our calibrations, in order to diminish the largest source
of uncertainty in the Li controversy. In fact, a reanaly-
sis of the Li spectroscopic data using the present tem-
perature scale leads to a Li plateau with A(Li) = 2.37
(Mel´endez & Ram´ırez 2004), a value that is closer to
that suggested by WMAP. A discrepancy, although much
smaller than that reported in previous works, still per-
sists.
The present work aims to a better definition of the
temperature vs. color relations, taking into account the
effect of the different chemical compositions (as measured
by the metallicity [Fe/H]) observed in the atmospheres
of F, G, and K stars. In the first part of this work (here-
after Part I), we derived the temperatures of about a
thousand stars with the IRFM. Combining these results
with photometric measurements of the sample stars we
now proceed to calibrate the Teff : color : [Fe/H] relations
in the following photometric systems (the corresponding
colors are given in parenthesis): U BV (B − V ), uvby
(b − y), Vilnius (Y − V , V − S), Geneva (B2 − V1, B2 − G,
t), Johnson-Cousins (V − RC , V − IC , RC − IC ), DDO
(42 − 45, 42 − 48), Johnson-2MASS (V − J2, V − H2,
V −K2), Tycho (BT −VT ), and Tycho-2MASS (VT −K2).
This paper revisits the widely used results of Alonso
et al. (1996, 1999; hereafter AAM), as well as our earlier
extensions (Mel´endez & Ram´ırez 2003, hereafter MR03;
Ram´ırez & Mel´endez 2004a, hereafter RM04a). In §2 the
characteristics of the sample adopted (better described
in Part I) are given along with the sources of the photom-
etry. The nature of the calibration formulae and the fits
to the data are described in §3. The empirical tempera-
ture scale is tabulated and tested in §4. The comparison
with other Teff scales is given in §5 and the conclusions
are summarized in §6.
2. THE SAMPLE, TEMPERATURES AND PHOTOMETRY
ADOPTED
Approximately 80% of the sample we used to calibrate
the color-Teff relations come from AAM work. The stel-
lar metallicities, however, have been assigned according
to the '2003 updated' Cayrel de Strobel et al. (2001) cat-
alog or the photometric calibration we derived in Part I,
which is also based on this catalog. The remaining stars
are from a sample of planet-hosting stars (Ram´ırez &
Mel´endez 2004b), and a selected sample of metal-poor
and metal-rich dwarf and giant stars. The latter allow
a better coverage of the regions below [Fe/H] ∼ −3.0
and above [Fe/H] ∼ 0.1, as well as the dwarf metal-poor
cool end (Teff ∼ 4500 K), with reliable input data. The
stars for which kinematical metallicities were adopted in
Part I were excluded from the calibrations. See Sect. 3.2
in Part I for details and references.
Effective temperatures were derived in Part I from
an implementation of the IRFM (see e.g. Blackwell
et al. 1980). There we showed that the IRFM tem-
peratures, which have a mean uncertainty of 1.3%, are
very well scaled to the direct ones (those derived from
angular diameter and bolometric flux measurements) for
[Fe/H] > −0.6, and 4000 K< Teff < 6500 K for dwarfs
or 3800 K< Teff < 5000 K for giants. For the rest of the
atmospheric parameters space, we still rely on the capa-
bility of the Kurucz models (those adopted in Part I for
the IRFM implementation) to reproduce the low blan-
keting effects in the infrared (> 1µm).
The temperatures we derived are not strictly consis-
tent with the whole set of metallicities adopted since
not all of the [Fe/H] values given in the literature were
derived using IRFM temperature scales, i.e. a redeter-
mination of the iron abundances with our temperature
scale would be needed in order to have a consistent set
of Teff and [Fe/H]. However, errors of 100 K in Teff re-
sult in errors of about 0.05-0.10 dex in [Fe/H] for both
dwarfs (e.g. Reddy et al. 2003, Gratton et al. 2003) and
giants (Shetrone 1996, Mel´endez et al. 2003, Francois
et al. 2003), which, in turn, may affect the Teff by only
about 10 K in a second iteration with the IRFM. There-
fore, these small inconsistencies do not significantly have
an impact on our IRFM calibrations. Furthermore, the
adoption of the mean of several metallicity determina-
tions as well as the use of hundreds of stars to define the
Teff : color : [Fe/H] relations minimize the effect.
The IRFM implementation from Part I uses essentially
only the V magnitude and the infrared photometry. The
bolometric fluxes were obtained from calibrations that
use only K and (V − K) and have an internal accuracy
of 1% whereas the systematic errors on the calibration
used, if present, will not affect the temperatures by more
than about a conservative estimate of 50 K (Sect. 3.4
in Part I, see also Ram´ırez & Mel´endez 2004b). Photo-
metric errors are easily propagated and are reflected in
the IRFM temperatures. Despite this, whenever reliable
photometry is adopted, the IRFM temperatures are ac-
curate to within ∼75 K. Since we do not expect this to
be the case for the whole sample, we strongly recommend
the use of several of the color-temperature calibrations
derived here, and give the mean value a larger weight
than the temperature from the IRFM, where available.
This reduces not only the errors in Teff introduced by
the photometry in the color-Teff calibrations, but also
the error due to the IRFM Teff.
U BV , uvby, Vilnius, Geneva, RI(Cousins), and DDO
photometry has been taken from several catalogs in-
cluded in the General Catalogue of Photometric Data
(Mermilliod et al. 1997, hereafter GCPD). Mean values
of (B − V ) and (b − y), as given in the GCPD, have
been adopted. Due to the low number of giants with
RI(Cousins) photometry available in the GCPD, we took
Washington or Kron-Eggen photometry and put them
The effective temperature scale of FGK stars. II.
3
into the Cousins system by means of the transformation
equations of Bessell (1979, 2001). This is the only place
where color-color transformations have been used, and so
the calibrations for giants in the Cousins system must be
taken with care. Note, however, that the filters involved
are not very different from those of the Cousins system,
especially for the Washington system. Photometry from
the Hipparcos-Tycho mission (ESA 1997) was also used,
as well as the infrared photometry from the final release
of the Two Micron All Sky Survey (2MASS, Cutri et al.
2003). We discuss each system in turn.
U BV (Johnson & Morgan 1953) and uvby (Stromgren
1966) were selected because they are widely used.
In
order to use existing photometry of thousands of stars
in other systems we also calibrated the visual and in-
frared colors in the Vilnius (Kararas et al. 1966, see
also Straizys & Sviderskiene 1972) and Cousins (1976)
systems, respectively.
Geneva photometry (Golay 1966) is considered one of
the major systems available. It is both very homogeneous
and the number of stars observed in both hemispheres is
large. We employed the t ≡ (B2 − G) − 0.39(B1 − B2)
parameter (Straizys 1995), whose metallicity sensitivity
is not as strong as in other Geneva colors, in addition to
the (B2 − V1) and (B2 − G) colors.
Although the colors from the DDO system (McClure &
van den Bergh 1968) are very metallicity dependent and
the number of stars observed is not particularly large,
the extremely careful observations performed with this
system allow the metallicity effects to be easily distin-
guished from photometric uncertainties, and so when-
ever reliable abundances (and DDO photometry) are
available, accurate temperatures may be derived. Both
C(42 − 45) and C(45 − 48) are satisfactory temperature
indicators (besides the strong metallicity dependence),
but the dispersion in the Teff vs. C(45 − 48) plane
is too large, and so we have preferred to calibrate the
C(42 − 48) ≡ C(42 − 45) + C(45 − 48) color.
A very large number of stars has been observed with
the all sky surveys of Hipparcos-Tycho and 2MASS. The
first of these contains mainly bright stars (VT < 12) while
the very bright star (V < 5) photometry of the second
one has lower quality and was not considered in general.
Note, however, that the 2MASS photometry is still ac-
curate for stars as faint as V ∼ 14, and those have been
included in the present work. 2MASS photometry was
adopted whenever the error in the K magnitude was less
than 0.025.
Reddening corrections for (B − V ) were already given
in Part I. The reddening ratios used to correct the re-
maining colors, k = E(color)/E(B − V ), are given in Ta-
ble 1. They were mainly obtained from Schlegel et al.
(1998) table of 'relative extinction for selected band-
passes,' adopting the appropriate effective wavelengths
of the filters, as given in the Asiago Database of Photo-
metric Systems (Fiorucci & Munari 2003, for r = 1.00
and E(B − V ) = 0). Reddening ratios k(color) ob-
tained in this way are in very good agreement with the
values given or calculated from related results in the
literature (Table 1). For the Johnson-Cousins colors,
however, interpolation from the Schlegel et al.
(1998)
tables is not a safe procedure given that the effective
wavelengths of the Cousins filters strongly depend on
spectral type (Bessell 1986). Using the effective wave-
lengths given by Fiorucci & Munari (2003) for RC and
IC we obtain k(V − RC ) = 0.51, k(V − IC ) = 1.12, and
k(RC − IC ) = 0.62 while the values often quoted in the
literature are around 0.6, 1.3, and 0.7, respectively (Dean
et al. 1978, Taylor 1986, Bessell et al. 1998). The latter
values are preferred.
3. THE CALIBRATIONS
The fits to the data were performed in a two-step pro-
cedure described as follows:
(1) All data points were iteratively fitted to
θeff = a0 + a1X + a2X 2 + a3X[Fe/H]
+a4[Fe/H] + a5[Fe/H]2 ,
(1)
where θeff = 5040/Teff, X represents the color, and ai
(i = 1, . . . , 5) the coefficients of the fit. In every iteration,
the points departing more than 2.5σ from the mean fit
were discarded. Normally, five to seven iterations were
required.
eff − T cal
The particular analytical expresion adopted (Eq. 1)
reasonably reproduces the observed trends, and it has
also some physical meaning (see e.g. Sect. 3 in RM04a).
(2) Whenever necessary and reasonable (see below),
the residuals of the fit (T IRFM
eff ) were fitted to poly-
nomials in X to remove any small systematic trends due
to the incapability of Eq. (1) to reproduce the effects of
spectral features such as the Balmer lines, the G band,
or the Paschen Jump on the observed colors. The poly-
nomial fits were performed in metallicity bins, and since
they rarely exceed 50 K, continuity is not severely com-
promised. The polynomial fits P (X, [Fe/H]) need to be
added to Eq. (1) so the final form is
Teff =
5040
θeff
+ P (X, [Fe/H]) .
(2)
Polynomial fits were only performed when enough stars
defined a clear trend in the residuals and care was taken
as not to force unphysical, artificial results. Neglecting
the polynomial fits in this procedure would lead to sys-
tematic errors of the order of 30 K or 40 K.
Due to the nature of the fits (particularly the poly-
nomial corrections), extrapolation leads to unreliable re-
sults. If necessary, one may, as a last resort, extrapolate
from the tables given in §4.
The coefficients of the fits (ai), the number of stars
included (N ), and the standard deviations (σ(Teff )) of
the seventeen color calibrations performed are given in
Tables 2 (dwarfs) and 3 (giants). The ranges of applica-
bility (Xmin < X < Xmax) in the following metallicity
bins: −0.5 < [Fe/H] < +0.5, −1.5 < [Fe/H] ≤ −0.5,
−2.5 < [Fe/H] ≤ −1.5, −3.5 < [Fe/H] ≤ −2.5; and
the coefficients of the polynomial fits (P = Pi PiX i)
are given in Tables 4 (dwarfs) and 5 (giants). Figs. 1-4
illustrate some of the calibrations in the Teff vs. color
planes, and the residuals of the fits (after the polyno-
mial corrections). The complete set of figures, for all the
calibrations, is available online.2 The figures illustrate,
better than the tables show, how far in [Fe/H] (both at
the metal-poor and metal-rich ends) a particular color
calibration may be applied.
The number of stars in the dwarf calibrations is al-
ways larger than 120. The standard deviations range
2 https://webspace.utexas.edu/ir68/teff
4
Ram´ırez & Mel´endez
Adopted Extinction Ratios and Comparison with the Literature
TABLE 1
Color
System
(b − y)
(Y − V )
(V − S)
(B2 − G)
(B2 − V1)
t
(V − RC )
(V − IC )
(RC − IC )
C(42 − 45)
C(42 − 48)
(V − J2)
(V − H2)
(V − K2)
(BT − VT )
(VT − K2)
Stromgren
Vilnius
Vilnius
Geneva
Geneva
Geneva
Johnson-Cousins
Johnson-Cousins
Cousins
DDO
DDO
Johnson-2MASS
Johnson-2MASS
Johnson-2MASS
Tycho
Tycho-2MASS
1k = E(color)/E(B − V ).
k1
0.74
0.72
0.62
1.14
0.86
0.98
0.60
1.30
0.70
0.23
0.58
2.16
2.51
2.70
1.02
2.87
k(Literature)
Reference
0.74
0.74
0.62
1.14
0.85
0.99
0.60
Crawford (1975)
Kuriliene & Sudzius (1974)
Sudzius et al. (1996)
Bersier (1996)
Bersier (1996)
Bersier (1996)
Taylor (1986)
1.25, 1.34
Dean et al. (1978), Taylor (1986)
0.70
0.23
0.59
2.25
2.55
2.72
· · ·
· · ·
Taylor (1986)
Dawson (1978)
Dawson (1978)
McCall (2004)
McCall (2004)
McCall (2004)
· · ·
· · ·
Coefficients of the Dwarf Star Color Calibrations1
TABLE 2
color (X)
a0
a1
a2
a3
a4
a5
N σ(Teff )
(B − V )
(b − y)
(Y − V )
(V − S)
(B2 − V1)
(B2 − G)
t
(V − RC )
(V − IC )
(RC − IC )
C(42 − 45)
C(42 − 48)
(BT − VT )
(V − J2)
(V − H2)
(V − K2)
(VT − K2)
0.5002
0.4129
0.0644
0.2417
0.6019
0.8399
0.7696
0.4333
0.3295
0.2919
0.5153
0.1601
0.5619
0.4050
0.4931
0.4942
0.4886
0.6440
1.2570
1.7517
1.3653
0.7663
0.4909
0.5927
1.4399
0.9516
2.1141
0.5963
0.4533
0.4462
0.4792
0.3056
0.2809
0.2773
-0.0690
-0.2268
-0.5264
-0.3823
-0.0713
-0.0666
0.3439
-0.5419
-0.2290
-1.0723
-0.0572
-0.0135
-0.0029
-0.0617
-0.0241
-0.0180
-0.0195
-0.0230
-0.0242
-0.0044
-0.0387
-0.0339
-0.0360
-0.0437
-0.0481
-0.0316
-0.0756
-0.0573
-0.0471
0.0003
-0.0392
-0.0396
-0.0294
-0.0300
-0.0566
-0.0464
-0.0407
-0.0105
-0.0382
-0.0468
-0.0143
-0.0239
0.0003
0.0267
-0.0221
0.0305
-0.0746
0.0401
0.0678
0.0444
0.0467
-0.0170
-0.0200
-0.0132
-0.0077
-0.0137
-0.0124
-0.0088
-0.0125
-0.0081
-0.0041
-0.0018
-0.0020
-0.0190
-0.0023
0.0020
-0.0008
-0.0008
495
434
159
142
358
368
308
133
127
137
120
133
378
361
364
397
318
1N is the number of stars employed and σ the standard deviation of each fit.
88
87
121
95
74
66
66
84
68
76
70
70
104
62
57
50
59
Coefficients of the Giant Star Color Calibrations1
TABLE 3
color (X)
a0
a1
a2
a3
a4
a5
N σ(Teff )
(B − V )
(b − y)
(Y − V )
(V − S)
(B2 − V1)
(B2 − G)
t
(V − RC )
(V − IC )
(RC − IC )
C(42 − 45)
C(42 − 48)
(BT − VT )
(V − J2)
(V − H2)
(V − K2)
(VT − K2)
0.5737
0.5515
0.3672
0.3481
0.6553
0.8492
0.7460
0.3849
0.3575
0.4351
0.4783
0.0023
0.5726
0.2943
0.4354
0.4405
0.4813
0.4882
0.9085
1.0467
1.1188
0.6278
0.4344
0.8151
1.6205
0.9069
1.6549
0.7748
0.6401
0.4461
0.5604
0.3405
0.3272
0.2871
-0.0149
-0.1494
-0.1995
-0.2068
-0.0629
-0.0365
-0.1943
-0.6395
-0.2025
-0.7215
-0.1361
-0.0632
-0.0324
-0.0677
-0.0263
-0.0252
-0.0203
0.0563
0.0616
0.0650
0.0299
0.0627
0.0466
0.0855
0.1060
0.0395
-0.0610
-0.0712
-0.0023
0.0518
0.0179
-0.0012
-0.0016
-0.0045
-0.1160
-0.0668
-0.0913
-0.0481
-0.0816
-0.0696
-0.0421
-0.0875
-0.0551
0.0332
-0.0117
-0.0706
-0.1170
-0.0532
-0.0049
-0.0053
0.0062
-0.0114
-0.0083
-0.0133
-0.0083
-0.0084
-0.0107
-0.0034
-0.0089
-0.0061
-0.0023
0.0071
-0.0070
-0.0094
-0.0088
-0.0027
-0.0040
-0.0019
269
208
159
152
200
189
192
90
95
128
188
191
261
163
177
182
112
1N is the number of stars employed and σ the standard deviation of each fit.
51
68
78
69
45
39
44
41
40
62
57
49
82
38
32
28
39
The effective temperature scale of FGK stars. II.
5
Ranges of Applicability per Metallicity Bin and Coefficients of the Polynomial Fits for the Dwarf
Star Calibrations
TABLE 4
color (X)
[Fe/H]1 Xmin Xmax
P0
P1
P2
P3
P4
P5
P6
(B − V )
(B − V )
(B − V )
(B − V )
(b − y)
(b − y)
(b − y)
(b − y)
(Y − V )
(Y − V )
(Y − V )
(Y − V )
(V − S)
(V − S)
(V − S)
(V − S)
(B2 − V1)
(B2 − V1)
(B2 − V1)
(B2 − V1)
(B2 − G)
(B2 − G)
(B2 − G)
(B2 − G)
t
t
t
t
(V − RC )
(V − RC )
(V − RC )
(V − RC )
(V − IC )
(V − IC )
(V − IC )
(RC − IC )
(RC − IC )
(RC − IC )
(RC − IC )
C(42 − 45)
C(42 − 45)
C(42 − 48)
C(42 − 48)
C(42 − 48)
(BT − VT )
(BT − VT )
(BT − VT )
(BT − VT )
(V − J2)
(V − J2)
(V − J2)
(V − J2)
(V − H2)
(V − H2)
(V − H2)
(V − H2)
(V − K2)
(V − K2)
(V − K2)
(V − K2)
(VT − K2)
(VT − K2)
(VT − K2)
(VT − K2)
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
+0.0
−1.0
−2.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
0.310
0.307
0.335
0.343
0.248
0.234
0.290
0.270
0.420
0.452
0.455
0.446
0.370
0.410
0.441
0.438
0.119
0.132
0.178
0.185
-0.271
-0.262
-0.200
-0.179
-0.119
-0.066
-0.006
0.020
0.204
0.284
0.264
0.240
0.491
0.597
0.547
0.242
0.300
0.283
0.290
0.461
0.480
1.286
1.465
1.399
0.344
0.391
0.380
0.367
0.815
0.860
0.927
0.891
0.839
1.032
1.070
1.093
0.896
1.060
1.101
1.126
0.942
1.078
1.237
1.170
1.507
1.202
1.030
0.976
0.824
0.692
0.672
0.479
0.940
0.660
0.720
0.643
1.130
0.690
0.810
0.584
0.936
0.593
0.621
0.435
1.110
0.502
0.544
0.150
0.450
0.373
0.333
0.295
0.880
0.546
0.532
0.336
1.721
1.052
1.026
0.838
0.718
0.551
0.364
1.428
0.812
2.711
1.957
1.509
1.715
1.556
0.922
0.504
2.608
2.087
1.983
1.932
3.215
2.532
2.535
2.388
3.360
2.665
2.670
2.596
3.284
2.561
2.406
1.668
-261.548
-324.033
30.5985
139.965
-1237.11
-2617.66
103.927
-294.106
-10407.1
11.6451
-507.732
-310.166
-1436.48
-728.818
101.031
596.461
-439.817
-257.527
-28.5544
64.2911
-6.29800
21.3254
11.5114
-156.547
-16.3530
35.2419
11.9635
-39.1918
-2666.55
4.20153
123.940
8.55498
-2757.79
-22.9008
-667.732
-3326.97
12.4740
-5837.31
32.1826
1533.40
808.065
658.568
176.678
1069.18
1199.21
-64.1045
-6030.19
-3255.07
422.406
-466.616
-862.072
-1046.10
-53.5574
1.60629
506.559
-471.588
-1425.36
2.35133
-1849.46
215.721
-1581.85
68.1279
-2384.82
-628.682
684.977
1516.44
-46.7882
-292.329
6591.29
22607.4
-312.419
648.320
42733.6
· · ·
1943.73
496.709
5566.00
2256.18
114.354
-1130.92
2637.06
2078.96
228.735
-365.124
160.976
-56.4562
-34.5752
-313.408
273.725
-185.953
· · ·
· · ·
27264.5
· · ·
-342.217
· · ·
9961.33
40.2078
1709.88
26263.8
· · ·
41439.2
· · ·
-5546.94
-2725.54
-283.310
-204.699
-678.907
-5470.57
140.575
29153.4
16259.9
-910.603
658.349
1236.84
1652.06
36.0990
· · ·
-1277.52
643.972
3218.36
· · ·
4577.00
-796.519
3273.10
-130.968
4196.14
423.682
-470.049
-2107.37
· · ·
· · ·
-11061.3
-68325.4
225.430
· · ·
-27378.8
· · ·
-1727.66
· · ·
-6780.53
-1704.54
-447.778
· · ·
-4762.80
-4919.04
-295.958
· · ·
-386.520
-651.533
-265.563
4886.53
-1383.02
· · ·
· · ·
· · ·
-103923.
· · ·
· · ·
· · ·
-10546.6
· · ·
-1069.62
-75355.8
· · ·
-94729.8
· · ·
6324.29
2806.13
-709.877
53.2421
· · ·
8367.46
-59.4233
-25882.7
-20315.3
621.335
-220.454
-423.729
-597.340
15.6878
· · ·
939.519
-199.639
-2566.54
· · ·
-4284.02
714.423
-2395.38
52.8391
-2557.04
· · ·
79.8977
852.150
· · ·
· · ·
5852.18
86072.5
· · ·
· · ·
-96466.3
· · ·
· · ·
· · ·
2613.40
· · ·
· · ·
· · ·
2606.79
3685.65
· · ·
· · ·
250.628
720.639
· · ·
· · ·
2274.81
· · ·
· · ·
· · ·
174663.
· · ·
· · ·
· · ·
-1746.05
· · ·
· · ·
94246.5
· · ·
69584.8
· · ·
-2254.52
-902.995
575.693
· · ·
· · ·
-5119.55
· · ·
-64112.9
· · ·
-132.566
· · ·
· · ·
· · ·
-8.84468
· · ·
-208.621
· · ·
859.644
· · ·
1770.38
-175.678
736.352
· · ·
595.365
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-38602.2
· · ·
· · ·
162033.
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-500.348
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-104940.
· · ·
· · ·
· · ·
10512.3
· · ·
· · ·
-43334.8
· · ·
· · ·
· · ·
· · ·
· · ·
-114.834
· · ·
· · ·
1078.09
· · ·
126115.
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-102.554
· · ·
-268.589
· · ·
-80.8177
· · ·
-31.9955
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-70956.4
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-23249.4
· · ·
· · ·
· · ·
-6653.57
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-59817.9
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
32644.9
· · ·
· · ·
· · ·
1301.21
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1Metallicity bins coded as: [Fe/H] = +0.0 (−0.5 < [Fe/H] < +0.5), [Fe/H] = −1.0 (−1.5 < [Fe/H] ≤ −0.5), [Fe/H] = −2.0
(−2.5 < [Fe/H] ≤ −1.5), [Fe/H] = −3.0 (−4.0 < [Fe/H] ≤ −2.5).
6
Ram´ırez & Mel´endez
Ranges of Applicability per Metallicity Bin and Coefficients of the Polynomial Fits for the Giant
Star Calibrations
TABLE 5
color (X)
[Fe/H]1 Xmin Xmax
P0
P1
P2
P3
P4
(B − V )
(B − V )
(B − V )
(B − V )
(b − y)
(b − y)
(b − y)
(b − y)
(Y − V )
(Y − V )
(Y − V )
(Y − V )
(V − S)
(V − S)
(V − S)
(V − S)
(B2 − V1)
(B2 − V1)
(B2 − V1)
(B2 − V1)
(B2 − G)
(B2 − G)
(B2 − G)
(B2 − G)
t
t
t
t
(V − RC )
(V − RC )
(V − RC )
(V − RC )
(V − IC )
(V − IC )
(V − IC )
(V − IC )
(RC − IC )
(RC − IC )
(RC − IC )
(RC − IC )
C(42 − 45)
C(42 − 45)
C(42 − 45)
C(42 − 45)
C(42 − 48)
C(42 − 48)
C(42 − 48)
C(42 − 48)
(BT − VT )
(BT − VT )
(BT − VT )
(BT − VT )
(V − J2)
(V − J2)
(V − J2)
(V − J2)
(V − H2)
(V − H2)
(V − H2)
(V − H2)
(V − K2)
(V − K2)
(V − K2)
(V − K2)
(VT − K2)
(VT − K2)
(VT − K2)
(VT − K2)
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
+0.0
−1.0
−2.0
−3.0
0.144
0.664
0.605
0.680
0.053
0.309
0.388
0.404
0.230
0.558
0.544
0.510
0.261
0.508
0.529
0.573
-0.079
0.385
0.307
0.407
-0.543
0.155
0.132
0.104
0.072
0.064
0.166
0.215
0.299
0.387
0.429
0.394
0.573
0.795
0.870
0.812
0.413
0.383
0.434
0.364
0.409
0.430
0.441
0.490
1.531
1.400
1.466
1.571
0.123
0.424
0.534
0.465
1.259
1.030
1.033
0.977
1.194
1.293
1.273
1.232
1.244
1.366
1.334
1.258
1.107
1.403
1.339
1.668
1.668
1.558
1.352
1.110
1.077
0.893
0.702
0.683
1.290
0.940
0.817
0.830
1.230
0.992
0.990
0.790
1.321
1.021
0.958
0.648
1.230
0.966
0.991
0.437
0.970
0.766
0.619
0.511
1.106
0.752
0.598
0.550
2.000
1.524
1.303
1.095
0.793
0.771
0.725
0.545
1.369
1.270
0.894
0.640
2.767
2.647
2.260
1.799
1.953
1.644
1.356
1.026
2.400
3.418
2.679
2.048
3.059
4.263
3.416
2.625
3.286
4.474
3.549
2.768
3.944
3.157
3.750
2.722
112.116
-12.9762
606.032
-9.26209
-124.159
888.088
1867.63
348.237
-308.851
-36.6533
3038.83
2685.88
-1605.54
187.841
10.1750
-14.2019
-15.0383
80.1344
323.889
1403.86
-0.52642
26.1904
9.87980
232.248
-46.1506
27.8739
67.1191
122.254
-8.51797
-10.7764
61.9821
27.9886
0.42933
-0.14180
9.31011
-23.0514
61.3557
-16.8645
32.0870
-15.6318
-68.3798
-0.62507
-40.0150
-314.177
1006.40
-6.92065
-113.222
566.914
346.881
196.416
938.789
1112.46
-122.595
-10.3848
4.18695
-67.7716
-377.022
71.7949
-27.4190
-46.2946
-72.6664
86.0358
-6.96153
-943.925
-37.2128
-193.512
-2.02136
8.06982
-372.622
· · ·
-1248.79
· · ·
553.827
-2879.23
-6657.49
-659.093
1241.57
383.901
-8668.15
-7433.07
9118.16
-270.092
· · ·
· · ·
50.8876
-147.055
-1031.06
-4866.09
10.4471
-89.2171
· · ·
-1452.43
-60.1641
-84.1166
-139.127
-394.604
15.6675
· · ·
-78.7382
-100.149
· · ·
· · ·
· · ·
· · ·
-116.711
· · ·
· · ·
· · ·
109.259
· · ·
35.6803
636.443
-549.012
· · ·
57.3030
-329.631
-1690.16
-372.164
-1919.98
-2717.81
76.4847
· · ·
13.8937
28.9202
334.733
-55.5383
20.7082
20.1061
36.5361
-65.4928
14.3298
1497.64
31.2900
166.183
· · ·
· · ·
67.1254
· · ·
627.453
· · ·
-490.703
2097.89
5784.81
· · ·
-1524.60
-458.085
6067.04
4991.81
-17672.6
· · ·
· · ·
· · ·
-32.3978
· · ·
795.024
4029.75
-7.53155
· · ·
· · ·
1848.07
643.522
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-34.4503
· · ·
· · ·
· · ·
-649.212
· · ·
· · ·
· · ·
2035.65
126.196
929.779
1577.18
· · ·
· · ·
· · ·
· · ·
-69.8093
9.61821
· · ·
· · ·
· · ·
10.8901
· · ·
-795.867
-6.72743
-33.2781
· · ·
· · ·
395.333
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
593.157
· · ·
· · ·
· · ·
14184.1
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-599.555
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
534.912
· · ·
· · ·
· · ·
-797.248
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
138.965
· · ·
· · ·
· · ·
· · ·
-203.471
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-4023.76
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-100.038
· · ·
· · ·
· · ·
70.7799
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
P5
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
P6
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1Metallicity bins coded as: [Fe/H] = +0.0 (−0.5 < [Fe/H] < +0.5), [Fe/H] = −1.0 (−1.5 < [Fe/H] ≤ −0.5), [Fe/H] = −2.0
(−2.5 < [Fe/H] ≤ −1.5), [Fe/H] = −3.0 (−4.0 < [Fe/H] ≤ −2.5).
The effective temperature scale of FGK stars. II.
7
7000
6000
5000
4000
0.3
0.6
0.9
1.2
1.5
Fig. 1. -- Left: Teff vs. (B − V ) observed for dwarfs in the metallicity bins: −0.5 < [Fe/H] ≤ +0.5 (filled circles), −1.5 < [Fe/H] ≤ −0.5
(open circles), −2.5 < [Fe/H] ≤ −1.5 (squares), and [Fe/H] ≤ −2.5 (triangles). The lines corresponding to our calibration for [Fe/H] = 0.0
− T cal
(solid line), [Fe/H] = −1.0 (dotted line), and [Fe/H] = −2.0 (dashed line) are also shown. Right: residuals of the fit (∆Teff = T IRFM
eff )
as a function of color and [Fe/H].
eff
7000
6000
5000
4000
1.0
2.0
3.0
Fig. 2. -- As in Fig. 1 for (V − K2) (dwarfs).
8
Ram´ırez & Mel´endez
8000
7000
6000
5000
4000
0.3
0.6
0.9
1.2
1.5
Fig. 3. -- As in Fig. 1 for (B − V ) (giants).
6000
5000
4000
2.0
3.0
4.0
Fig. 4. -- As in Fig. 1 for (V − K2) (giants).
from 50 K (for V − K2) to 121 K (for Y − V ). For giants
the number of stars in the calibrations amount from 90 to
270, while the standard deviations range from 28 K (for
V − K2) to 82 K (for BT − VT ). The standard deviations
are in general lower than those obtained by AAM, which
is a consequence of adopting more accurate input param-
eters. For the (B − V ) calibration, for example, AAM
obtained dispersions of 130 K for dwarfs and 100 K for
giants. Our values are 88 K and 51 K, respectively. The
general trends in the Teff vs. color planes in common
with AAM, however, are very similar (see §5).
4. THE EMPIRICAL IRFM TEMPERATURE SCALE
4.1. Intrinsic colors of dwarfs and giants
By numerically inverting Eq. (2) for a given Teff, [Fe/H]
pair, grids of colors in the (Teff , [Fe/H]) space were de-
rived. This procedure is safer and easier than calibrating
the relations from the data, since it guarantees a sin-
gle correspondence between the three quantities involved,
i.e. the grids and the calibration formulae produce the
same results. The intrinsic colors of both dwarfs and gi-
ants as a function of Teff and [Fe/H] are given in Tables
6-11.
The effective temperature scale of FGK stars. II.
9
Intrinsic (B − V ), (b − y) and (Y − V ) Colors
TABLE 6
Teff / [Fe/H] = +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0
(B − V )
(b − y)
(Y − V )
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
7250
7500
7750
8000
8250
1.383
1.236
1.104
0.986
0.882
0.788
0.703
0.627
0.558
0.497
0.440
0.388
0.341
1.629
1.481
1.324
1.171
1.032
0.909
0.802
0.707
0.623
0.547
0.478
0.416
0.358
0.305
0.256
0.210
0.166
· · ·
· · ·
1.196
1.055
0.942
0.844
0.757
0.678
0.605
0.538
0.475
0.418
0.365
0.316
· · ·
· · ·
1.405
1.218
1.053
0.908
0.778
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.026
0.911
0.811
0.721
0.642
0.570
0.507
0.448
0.396
0.348
· · ·
· · ·
· · ·
· · ·
1.198
0.987
0.830
0.701
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Dwarf Stars
0.768
0.694
0.632
0.577
0.526
0.480
0.437
0.398
0.360
0.326
0.294
0.264
· · ·
· · ·
0.647
0.577
0.524
0.479
0.439
0.402
0.369
0.337
0.306
0.276
0.246
· · ·
Giant Stars
1.016
0.906
0.806
0.716
0.634
0.561
0.497
0.437
0.385
0.338
0.295
0.255
0.220
0.187
0.157
0.130
0.104
0.080
0.058
· · ·
· · ·
0.750
0.648
0.571
0.507
0.453
0.404
0.360
0.320
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.975
0.880
0.795
0.718
0.649
0.586
0.528
0.475
0.425
0.381
· · ·
· · ·
· · ·
· · ·
· · ·
0.980
0.780
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.656
0.593
0.536
0.489
0.445
0.406
0.371
0.340
0.311
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.569
0.497
0.442
0.397
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.442
0.403
0.370
0.339
0.311
0.286
· · ·
· · ·
· · ·
· · ·
0.674
0.590
0.515
0.446
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.925
0.861
0.785
0.724
0.674
0.632
0.595
0.559
0.526
0.493
0.461
0.427
· · ·
1.223
1.072
0.957
0.863
0.782
0.712
0.649
0.594
0.544
0.500
0.459
0.423
0.389
0.358
0.330
0.304
0.280
0.258
0.237
· · ·
· · ·
· · ·
· · ·
0.650
0.605
0.566
0.533
0.502
0.475
· · ·
· · ·
· · ·
· · ·
· · ·
0.902
0.815
0.737
0.667
0.603
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.695
0.649
0.606
0.568
0.533
0.503
0.474
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.712
0.642
0.589
0.544
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.620
0.579
0.542
0.511
0.483
0.458
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.742
0.664
0.605
0.556
0.513
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
4.2. The IRFM vs. direct temperature scales
In Part I we compared the IRFM temperatures with
the direct temperatures for both dwarf and giant stars.
The comparison showed that the zero points of both tem-
perature scales are essentially equal. As a practical ex-
ample of the use of the present calibrations, and to test
their reliability, here we derive the temperatures from
the colors of the same stars used for the comparison of
direct and IRFM temperatures in Part I.
eff
In Table 12 we give the temperatures derived from an-
gular diameter and bolometric flux measurements (T dir
eff ),
as given in Part I for dwarfs (see Sect. 5.1 in Part I for
the references) and by Mozurkewich et al. (2003) for gi-
ants, along with the temperatures we obtained with the
IRFM in Part I (T IRFM
). The average of the temper-
atures obtained from the color calibrations are given as
T cal
eff . The number of colors used (N ) for each star is also
provided. More than eight color calibrations were used.
As we showed in Part I, the photometric errors are im-
portant in our IRFM implementation. The IRFM tem-
peratures we obtained in Part I depend on the quality of
the V and infrared magnitudes, which, in general (but
not always), are reasonably accurate. With the color cal-
ibrations, the impact of photometric errors is removed as
an average Teff vs. color relation is constructed. When
using a given color calibration, the systematic error in the
obtained Teff is mainly due to the error in that particular
color, and only to a lesser extent due to the errors in the
IRFM temperatures. If more than one color is used, the
photometric errors from different colors may be reduced
by taking an average Teff. Consequently, if an IRFM
temperature is available for a given star, and a mean Teff
is obtained from its colors, the photometric temperature,
provided that the number of colors used is large enough,
will be more reliable. We have already shown this in the
case of Arcturus (see Sect. 3.5 in RM04a).
Thus, whenever an IRFM temperature is available, we
suggest the following temperature be adopted:
Teff =
mT cal
eff + nT IRFM
eff
m + n
,
(3)
where m ≥ n. The exact values of m and n should be
chosen depending on the quality of the colors and the
IRFM temperature. The values we have used for the
Teff (adopted) in Table 12 are m = 2, n = 1.
The mean difference T cal
is −16 ± 48 K for
dwarfs and −6 ± 37 K for giants. Fig. 5 shows that
no systematic errors are introduced by the calibrations,
neither with Teff nor [Fe/H].
eff − T IRFM
eff
Given the reliability of the photometric temperatures,
they may be safely combined with the IRFM tempera-
10
Ram´ırez & Mel´endez
Intrinsic (V − S), (B2 − V1) and (B2 − G) Colors
TABLE 7
Teff / [Fe/H] = +0.0 −1.0 −2.0 −3.0
+0.0
(V − S)
(B2 − V1)
−1.0 −2.0 −3.0
(B2 − G)
+0.0
−1.0
−2.0
−3.0
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
7250
7500
7750
8000
8250
1.038
0.928
0.842
0.770
0.705
0.648
0.597
0.551
0.509
0.472
0.437
0.407
0.379
1.147
0.999
0.892
0.809
0.739
0.679
0.625
0.577
0.532
0.491
0.453
0.418
0.384
0.353
0.324
0.297
0.271
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.668
0.614
0.566
0.523
0.485
0.450
0.419
· · ·
· · ·
· · ·
0.953
0.862
0.783
0.714
0.652
0.598
0.548
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.755
0.698
0.647
0.602
0.560
0.524
0.490
0.459
· · ·
· · ·
· · ·
· · ·
· · ·
0.888
0.795
0.716
0.647
0.587
0.535
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.558
0.525
0.496
0.469
0.444
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.728
0.658
0.596
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Dwarf Stars
· · ·
0.806
0.711
0.631
0.559
0.494
0.432
0.374
0.320
0.269
0.221
0.176
0.134
· · ·
· · ·
· · ·
0.561
0.494
0.434
0.378
0.326
0.277
0.230
0.187
0.146
· · ·
Giant Stars
1.253
1.081
0.934
0.807
0.696
0.599
0.512
0.435
0.365
0.303
0.246
0.194
0.146
0.102
0.062
0.025
-0.010
-0.042
-0.073
· · ·
1.017
0.870
0.741
0.627
0.526
0.436
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.553
0.484
0.422
0.365
0.314
0.267
0.224
0.186
· · ·
· · ·
· · ·
· · ·
0.941
0.731
0.590
0.478
0.384
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.419
0.370
0.325
0.284
0.246
0.212
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.554
0.457
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.005
0.791
0.626
0.487
0.365
0.259
0.163
0.076
-0.002
-0.073
-0.138
-0.197
-0.251
· · ·
1.038
0.835
0.662
0.511
0.378
0.262
0.156
0.063
-0.022
-0.099
-0.169
-0.233
-0.292
-0.346
-0.396
-0.443
-0.486
-0.526
· · ·
· · ·
0.460
0.350
0.253
0.165
0.085
0.010
-0.059
-0.123
-0.182
-0.238
· · ·
· · ·
0.530
0.418
0.316
0.223
0.138
0.060
-0.010
-0.074
-0.133
-0.187
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.117
0.037
-0.023
-0.072
-0.115
-0.152
· · ·
· · ·
· · ·
0.944
0.739
0.560
0.404
0.265
· · ·
· · ·
0.777
0.550
0.358
0.192
· · ·
· · ·
· · ·
· · ·
0.318
0.181
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Fig. 5. -- Difference between the temperatures from the color cal-
ibrations and the IRFM temperatures as a function of the adopted
temperatures and metallicities of dwarfs (filled circles) and giants
(open circles).
Fig. 6. -- Difference between the adopted temperatures and
direct temperatures as a function of the adopted temperatures and
metallicities of dwarfs (filled circles) and giants (open circles).
tures according to Eq. (3). When comparing the adopted
temperatures (those obtained by combining the photo-
metric and IRFM temperatures) with the direct ones, a
mean difference of −1 ± 60 K for dwarfs and −11 ± 50 K
for giants is obtained. As shown in Fig. 6, no trends are
observed with either Teff or [Fe/H]. The dispersion in
the mean differences for dwarfs is the same as the one
found in Part I when only the IRFM temperatures were
compared with the direct temperatures. However, when
we consider the adopted temperatures, the zero point of
the temperature scale is only 1 K below the direct one (it
was 30 K when only the IRFM temperatures were used
in Part I). For giants, the zero points are still in excellent
The effective temperature scale of FGK stars. II.
11
Intrinsic t, (V − RC ) and (V − IC ) Colors
TABLE 8
Teff / [Fe/H] = +0.0
−1.0
−2.0 −3.0 +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0
t
(V − RC )
(V − IC )
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
· · ·
· · ·
· · ·
· · ·
0.403
0.338
0.277
0.220
0.165
0.112
0.059
0.007
-0.046
-0.100
0.926
0.781
0.647
0.530
0.430
0.346
0.275
0.213
0.158
0.110
· · ·
· · ·
· · ·
· · ·
0.366
0.308
0.252
0.200
0.149
0.100
0.053
0.007
-0.038
· · ·
· · ·
· · ·
0.652
0.528
0.424
0.335
0.259
0.193
0.134
0.082
· · ·
· · ·
· · ·
· · ·
· · ·
0.314
0.259
0.206
0.157
0.109
0.064
0.020
· · ·
· · ·
· · ·
· · ·
· · ·
0.561
0.440
0.337
0.250
0.175
· · ·
· · ·
Dwarf Stars
· · ·
0.822
0.756
0.674
0.579
0.498
0.440
0.393
0.355
0.321
0.290
0.261
0.233
0.206
· · ·
· · ·
· · ·
· · ·
0.509
0.454
0.407
0.366
0.330
0.298
· · ·
· · ·
· · ·
· · ·
Giant Stars
0.948
0.782
0.673
0.592
0.527
0.472
0.426
0.387
0.351
0.321
· · ·
· · ·
0.657
0.568
0.498
0.440
0.391
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.269
0.220
0.174
0.131
0.090
0.051
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.439
0.332
0.241
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.488
0.440
0.399
0.362
0.329
0.300
0.273
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.577
0.497
0.432
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.318
0.291
0.267
0.245
· · ·
· · ·
· · ·
· · ·
· · ·
0.496
0.424
· · ·
· · ·
· · ·
· · ·
1.713
1.581
1.421
1.225
1.059
0.937
0.840
0.760
0.690
0.626
0.569
0.515
· · ·
· · ·
1.861
1.492
1.278
1.122
0.998
0.897
0.811
0.738
0.673
0.617
· · ·
· · ·
· · ·
· · ·
0.980
0.882
0.800
0.731
0.670
0.616
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.281
1.110
0.977
0.870
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.958
0.876
0.801
0.735
0.674
0.619
0.571
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.135
0.987
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.006
0.877
· · ·
· · ·
· · ·
· · ·
Intrinsic (RC − IC ), C(42 − 45) and C(42 − 48) Colors
TABLE 9
Teff / [Fe/H] = +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0
(RC − IC )
C(42 − 45)
C(42 − 48)
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
0.826
0.745
0.632
0.549
0.489
0.442
0.402
0.366
0.333
0.305
0.277
0.252
· · ·
· · ·
· · ·
0.720
0.618
0.541
0.479
0.427
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.693
0.600
0.532
0.479
0.435
0.397
0.366
0.338
0.314
· · ·
· · ·
· · ·
· · ·
· · ·
0.711
0.614
0.540
0.480
0.431
0.388
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.505
0.468
0.436
0.406
0.378
0.350
0.322
0.295
· · ·
· · ·
· · ·
· · ·
· · ·
0.632
0.560
0.501
0.451
· · ·
· · ·
· · ·
· · ·
· · ·
Dwarf Stars
· · ·
1.418
1.313
1.187
1.047
0.915
0.804
0.714
0.641
0.577
0.523
0.475
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.761
0.681
0.611
0.550
0.495
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Giant Stars
· · ·
1.322
1.149
1.009
0.892
0.793
0.708
0.633
0.567
0.508
0.455
1.253
1.093
0.962
0.853
0.758
0.677
0.605
0.542
0.486
0.435
· · ·
· · ·
· · ·
0.808
0.715
0.634
0.563
0.502
0.446
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.362
0.339
0.317
0.299
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.504
0.458
0.418
0.384
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.624
0.534
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.620
2.461
2.298
2.141
1.997
1.871
1.759
1.660
1.571
1.491
1.418
1.352
1.291
· · ·
2.676
2.456
2.252
2.080
1.938
1.817
1.715
1.627
1.547
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.843
1.738
1.644
1.558
1.480
· · ·
· · ·
· · ·
· · ·
· · ·
2.441
2.229
2.052
1.903
1.774
1.662
1.562
1.474
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.504
1.442
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.093
1.917
1.769
1.643
1.533
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.688
1.583
· · ·
· · ·
· · ·
· · ·
· · ·
12
Ram´ırez & Mel´endez
Intrinsic (BT − VT ), (V − J2) and (V − H2) Colors
TABLE 10
Teff / [Fe/H] = +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0
(BT − VT )
(V − J2)
(V − H2)
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
7000
7250
7500
7750
8000
1.563
1.427
1.295
1.165
1.035
0.910
0.797
0.698
0.613
0.541
0.478
0.422
0.373
1.934
1.773
1.592
1.400
1.213
1.045
0.904
0.784
0.682
0.595
0.518
0.450
0.389
0.334
0.284
0.237
0.194
0.154
1.464
1.297
1.145
1.009
0.886
0.775
0.674
0.582
0.498
0.420
· · ·
· · ·
· · ·
· · ·
1.630
1.386
1.190
1.022
0.878
0.750
0.637
0.534
0.441
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.912
0.825
0.704
0.622
0.554
0.486
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.106
0.906
0.761
0.645
0.543
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Dwarf Stars
· · ·
2.322
2.040
1.793
1.587
1.417
1.275
1.154
1.048
0.956
0.875
· · ·
· · ·
· · ·
2.072
1.875
1.695
1.534
1.390
1.262
1.147
1.045
0.955
0.873
· · ·
· · ·
Giant Stars
· · ·
· · ·
2.196
1.939
1.736
1.571
1.431
1.311
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.886
2.428
2.114
1.874
1.679
1.516
1.379
1.260
1.155
1.063
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.474
0.426
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.795
0.641
0.525
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.950
1.795
1.647
1.507
1.377
1.259
1.153
1.057
0.970
· · ·
· · ·
· · ·
· · ·
2.607
2.218
1.938
1.719
1.541
1.392
1.263
1.151
1.052
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.886
1.758
1.633
1.512
1.398
1.291
1.192
1.102
1.020
0.945
· · ·
· · ·
· · ·
· · ·
· · ·
2.001
1.763
1.572
1.412
1.278
1.160
1.058
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.899
2.562
2.267
2.010
1.784
1.587
1.414
1.260
1.121
0.998
0.886
· · ·
· · ·
· · ·
2.832
2.514
2.239
1.997
1.785
1.600
1.436
1.292
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.413
2.163
1.947
1.760
1.594
1.448
1.317
1.199
1.092
· · ·
· · ·
3.706
3.173
2.775
2.456
2.191
1.966
1.772
1.602
1.451
1.317
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.469
2.290
2.104
1.922
1.751
1.595
1.459
1.337
1.229
1.134
· · ·
· · ·
· · ·
3.257
2.830
2.493
2.215
1.981
1.780
1.606
1.451
1.315
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.342
2.172
2.013
1.863
1.725
1.597
1.480
1.372
1.275
1.184
1.101
· · ·
· · ·
· · ·
· · ·
2.494
2.221
1.990
1.791
1.618
1.466
1.330
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Intrinsic (V − K2) and (VT − K2) Colors
TABLE 11
Teff / [Fe/H] = +0.0 −1.0 −2.0 −3.0 +0.0 −1.0 −2.0 −3.0
(V − K2)
(VT − K2)
4250
4500
4750
5000
5250
5500
5750
6000
6250
6500
6750
3750
4000
4250
4500
4750
5000
5250
5500
5750
6000
6250
Dwarf Stars
2.609
2.437
2.234
2.022
1.829
1.662
1.515
1.386
1.268
1.161
· · ·
2.516
2.352
2.182
2.012
1.847
1.693
1.553
1.426
1.312
1.210
· · ·
Giant Stars
· · ·
3.436
2.978
2.617
2.323
2.076
1.863
1.679
1.518
1.374
· · ·
· · ·
· · ·
· · ·
2.660
2.328
2.085
1.883
1.706
1.547
1.401
1.268
· · ·
2.538
2.275
2.046
1.845
1.669
1.511
1.370
1.243
1.126
· · ·
3.911
3.322
2.893
2.554
2.274
2.039
1.834
1.657
1.500
· · ·
· · ·
3.106
2.728
2.374
2.097
1.874
1.682
1.512
1.357
1.212
1.074
0.942
· · ·
· · ·
3.005
2.625
2.318
2.064
1.847
1.660
1.497
1.351
· · ·
3.284
2.850
2.476
2.189
1.957
1.758
1.580
1.415
1.261
1.115
0.974
· · ·
3.626
3.148
2.759
2.436
2.164
1.928
1.723
1.542
1.380
1.236
· · ·
· · ·
2.441
2.162
1.932
1.738
1.568
1.420
1.288
1.169
· · ·
· · ·
· · ·
3.122
2.754
2.440
2.168
1.930
1.723
1.539
· · ·
· · ·
· · ·
· · ·
2.402
2.108
1.906
1.738
1.587
1.447
1.318
· · ·
· · ·
· · ·
3.565
3.106
2.734
2.427
2.166
1.940
1.744
1.570
1.415
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1.620
1.470
1.340
1.227
· · ·
· · ·
· · ·
· · ·
· · ·
2.462
2.205
1.981
1.786
· · ·
· · ·
· · ·
The effective temperature scale of FGK stars. II.
13
Comparison of Direct and IRFM Temperatures1
TABLE 12
HD
[Fe/H]
T dir
eff
T IRFM
eff
T cal
eff
N Teff (adopted)
10700
16160
22049
26965
61421
88230
121370
128620
128621
131977
198149
209100
209458
3546
3627
3712
6860
9927
10380
12533
12929
18884
29139
62509
62721
76294
80493
94264
99998
102224
113226
124897
135722
150997
164058
169414
181276
189319
197989
210745
214868
217906
221115
222107
Dwarf and Subgiant Stars
-0.54
-0.03
-0.12
-0.28
-0.01
-0.12
0.25
0.20
0.20
0.09
-0.18
-0.02
-0.01
-0.72
0.17
-0.10
-0.07
-0.01
-0.27
-0.07
-0.25
0.00
-0.18
-0.02
-0.27
-0.01
-0.26
-0.20
-0.39
-0.44
0.11
-0.55
-0.40
-0.28
-0.15
-0.16
0.02
0.00
-0.12
0.25
-0.25
-0.11
0.04
-0.50
5319 ± 43
· · ·
5078 ± 40
· · ·
6532 ± 39
3967 ± 63
6081 ± 47
5771 ± 23
5178 ± 23
4469 ± 57
4939 ± 41
4527 ± 29
· · ·
· · ·
4392 ± 27
4602 ± 29
· · ·
· · ·
· · ·
4254 ± 27
4493 ± 28
· · ·
3871 ± 24
4858 ± 30
· · ·
· · ·
3836 ± 24
· · ·
· · ·
· · ·
4981 ± 31
4226 ± 29
4851 ± 32
4841 ± 36
4013 ± 30
· · ·
· · ·
3858 ± 24
4757 ± 30
4351 ± 27
· · ·
· · ·
· · ·
· · ·
5372 ± 65
4714 ± 67
5015 ± 56
5068 ± 63
6591 ± 73
3950 ± 161
6038 ± 75
5759 ± 70
5201 ± 65
4571 ± 52
4907 ± 54
4642 ± 54
5993 ± 71
5270 ± 82
4781 ± 42
4981 ± 38
5099 ± 25
6586 ± 46
3985 ± 73
6034 ± 47
5736 ± 19
5103 ± 20
4545 ± 26
4903 ± 94
4605 ± 20
5983 ± 37
Giant Stars
4935 ± 52
4343 ± 45
4553 ± 48
3824 ± 41
4380 ± 48
4132 ± 46
4259 ± 45
4501 ± 50
3718 ± 46
3883 ± 44
4833 ± 50
3988 ± 48
4817 ± 50
3851 ± 42
4670 ± 51
3919 ± 45
4378 ± 46
5049 ± 59
4231 ± 49
4834 ± 50
4948 ± 54
3927 ± 42
4450 ± 50
4935 ± 54
3877 ± 41
4710 ± 52
4482 ± 51
4303 ± 47
3648 ± 43
4980 ± 63
4605 ± 49
4880 ± 44
4350 ± 20
4516 ± 51
3845 ± 45
4298 ± 60
4146 ± 19
4195 ± 50
4500 ± 25
3739 ± 31
3901 ± 33
4822 ± 20
4005 ± 32
4805 ± 83
3868 ± 14
4650 ± 56
3896 ± 10
4396 ± 26
4984 ± 50
4283 ± 42
4793 ± 27
4922 ± 34
3920 ± 21
4462 ± 26
4942 ± 21
3859 ± 15
4716 ± 20
4222 ± 144
4265 ± 25
3741 ± 132
4955 ± 20
4650 ± 34
10
12
11
10
9
9
8
4
6
5
10
11
6
10
8
7
9
8
13
4
10
7
10
8
5
11
7
10
8
8
10
13
10
9
10
7
10
6
10
9
7
8
9
10
5304 ± 105
4759 ± 79
4992 ± 68
5089 ± 68
6588 ± 86
3973 ± 177
6035 ± 88
5744 ± 72
5136 ± 68
4554 ± 58
4904 ± 108
4617 ± 58
5986 ± 80
4898 ± 68
4348 ± 49
4528 ± 70
3838 ± 61
4325 ± 77
4141 ± 50
4216 ± 67
4500 ± 56
3732 ± 55
3895 ± 55
4826 ± 54
3999 ± 58
4809 ± 97
3862 ± 44
4657 ± 76
3904 ± 46
4390 ± 53
5006 ± 77
4266 ± 65
4807 ± 57
4931 ± 64
3922 ± 47
4458 ± 56
4940 ± 58
3865 ± 44
4714 ± 56
4309 ± 153
4278 ± 53
3710 ± 139
4963 ± 66
4635 ± 60
1Metallicities, direct, and IRFM temperatures as given in Part I. For dwarfs only the
direct temperatures obtained with reliable angular diameter measurements are given;
for giants the direct temperatures (Teff > 3800 K only) are from Mozurkewich et al.
(2003). The temperatures obtained from N color calibrations are also given. The last
column corresponds to the suggested Teff to adopt: Teff = (2T cal
eff + T IRFM
eff
)/3.
agreement (at the 10 K level), and the dispersion in the
mean differences has been reduced by 10 K.
4.3. The colors of the Sun
Interpolation from Tables 6-11 at Teff = 5777 K and
[Fe/H] = 0.0 allowed us to derive colors representative of
a solar-twin star. They are given in Table 13 along with
the colors of five solar analogs from the list of Soubiran
& Triaud (2004). The metallicities given by Soubiran
& Triaud have been adopted to derive the photometric
temperatures of these stars, which are also given in Ta-
ble 13.
The 'closest ever solar twin', 18 Sco or HD 146233
(Porto de Mello & da Silva 1997) appears to be ∼ 40 K
cooler than the Sun. The other four stars in Table 13
are those from the list of 'Top Ten solar analogs in the
ELODIE library' (Soubiran & Triaud 2004) whose tem-
peratures are around 5780 K. The remaining five stars of
the 'Top Ten' are in general cooler by about 100 K.
The range of (B − V ) colors for these five solar analogs
is 0.61 < (B − V ) < 0.65, while the IRFM tempera-
ture scale suggests (B − V )⊙ = 0.62, implying reason-
able agreement considering photometric errors (which
are around 0.01 mag) and metallicity effects.
In gen-
eral, the remaining colors of the solar analogs are also
consistent with those we derived for the Sun with the
14
Ram´ırez & Mel´endez
Colors of the Sun and Solar Analogs
TABLE 13
color
Sun
HD 146233 HD 10307 HD 47309 HD 95128 HD 71148
(B − V )
(b − y)
(Y − V )
(V − S)
(B2 − V1)
(B2 − G)
t
(V − RC )
(V − IC )
(RC − IC )
C(42 − 45)
C(42 − 48)
(BT − VT )
(V − J2)
(V − H2)
(V − K2)
(VT − K2)
1
Teff
[Fe/H]2
0.619
0.394
0.556
0.546
0.368
0.067
0.159
0.351
0.682
0.330
0.633
1.650
0.689
1.141
1.396
1.495
1.562
5777 ± 10
+0.00
0.651
0.401
0.570
0.540
0.385
0.089
0.170
0.353
0.688
0.335
0.651
1.671
0.736
· · ·
· · ·
· · ·
· · ·
5735 ± 39
+0.05
0.616
0.389
0.540
0.550
0.362
0.063
0.153
· · ·
· · ·
· · ·
0.630
1.640
0.711
1.188
1.482
1.552
1.617
5796 ± 36
−0.02
0.623
0.412
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.686
· · ·
· · ·
· · ·
· · ·
5732 ± 47
+0.11
0.606
0.391
0.550
0.540
0.363
0.057
0.149
· · ·
· · ·
· · ·
0.629
1.637
0.733
1.191
1.433
1.508
1.578
5807 ± 15
+0.0
0.625
0.399
· · ·
· · ·
0.367
0.070
0.159
· · ·
· · ·
· · ·
0.634
1.647
0.692
1.171
1.451
1.499
1.563
5754 ± 27
−0.02
1The temperature of the Sun is the direct one, for the solar analogs the average of the tem-
peratures from colors are given. Simple standard deviations are given as error bars.
2[Fe/H] = +0.0 for the Sun, by definition. The metallicities of the solar analogs are from
Soubiran & Triaud (2004).
IRFM Teff scale, within photometric uncertainties.
As noted by Sekiguchi & Fukugita (2000), the values
of (B − V )⊙ most often found in the literature range
from 0.65 to 0.67, which are considerably larger than the
present result.
In their detailed study of the (B − V )
color-temperature relations they also find a bluer (B −
V )⊙ = 0.626. Therefore, in order to correctly place the
Sun in the Teff vs. (B − V ) plane, as defined by normal
stars, (B − V )⊙ should be bluer than what is usually
quoted in the literature.
4.4. Metallicity effects
Metallicity has an important effect on the Teff vs.
color relations, particularly for the (B − V ) color (e.g.
Cameron 1985, Mart´ınez-Roger et al.
1992). Fig. 7
shows representative theoretical spectra of metal-poor
and metal-rich dwarfs and giants at 4500 K and 6750 K.
The transmission functions of the filters adopted in the
present study are also shown. The figure is intended to
be of use when trying to understand the effects of both
[Fe/H] and log g on the Teff vs. color relations.
In general, at a given temperature, the colors get red-
der (larger) as more metals are present. The simplest
explanation is that with more metals in the atmosphere,
the UV and blue continuum is greatly reduced by line
blanketing, with a corresponding increase of the red con-
tinuum due to flux redistribution, which results in the
metal-rich stars being redder. This explanation is ap-
propriate for colors constructed with filters measuring
the fluxes in the "blue" (4000A< λ < 5500A) and "vi-
sual" (4500A< λ < 6000A) regions of the spectrum but
must not be extended to other colors straightforwardly.
The situation for the Cousins and Johnson-2MASS col-
ors, for example, is substantially different.
As it can be clearly seen in Fig. 2 (dwarfs) and to a
lesser extent in Fig. 4 (giants), the (V − K2) colors of
smaller) than
metal-poor stars are indeed 'bluer' (i.e.
the metal-rich stars at cool temperatures. However, for
temperatures above 6000 K the metal-poor stars are red-
der. This is consistent with the synthetic spectra shown
in Fig. 7. For Teff = 6750 K and log g = 4.5, the flux in
the K2 band does not change substantially as [Fe/H] is
reduced from +0.0 to −2.5 dex. The flux in the visual
region, on the other hand, is larger for the metal-rich
spectrum due to the redistribution of the UV flux into
the visual region. The net result is a larger (V − K2) for
the metal-poor stars (assuming these theoretical spectra
reasonably reproduce the real ones).
Given its relatively low dependence on [Fe/H], (V −K)
is a very good temperature indicator. We caution, how-
ever, that according to our results this is valid only from
4800 K to 6000 K for main sequence stars. If [Fe/H] is
unknown, one may be tempted to use the calibrations
for (V − K) assuming [Fe/H] = +0.0 and adopt a solar
metallicity Teff . If the temperature of the star results in
6000 K or more, however, and the star is metal-poor, for
example [Fe/H] ∼ −3.0, the temperature may be being
underestimated by 200 K or even more. Thus, at these
high temperatures, adopting a solar metallicity temper-
ature is unacceptable for metal deficient dwarfs. Results
from model atmospheres support this conclusion. The
situation for cool dwarfs (Teff < 4500 K) should not be
considered conclusive, because although it is consistent
with the models, the corresponding models themselves
are not very reliable. For giants the effect of a wrongly
adopted solar metallicity on Teff, for the very metal-poor
stars, is ∼100 K or less.
Ryan et al.
(1999) criticized Alonso et al.
(1996)
(b − y) calibration for dwarfs, arguing that it becomes
unphysical below [Fe/H] = −2.5 for turn-off stars (Teff ∼
6750 K), and attributed the effect to the quadratic [Fe/H]
term of the calibration formula (Eq. 1). Unphysical or
not, it is the IRFM that produces higher temperatures
for very metal-poor turn-off stars, an effect that is clearly
The effective temperature scale of FGK stars. II.
15
Fig. 7. -- Spectral Energy Distributions (SEDs) from Kurucz models, as given by Lejeune et al. (1997), for atmospheric parameters
representative of cool (Teff = 4500 K), hot (Teff = 6750 K), giant (log g = 2.5), dwarf (log g = 4.5), metal-poor ([Fe/H] = −2.5, dotted
lines), and metal-rich ([Fe/H] = +0.0, solid lines) stars. The filter transmission functions of interest to this work are also shown.
seen not only in the Teff vs. (b − y) plane but also in the
Teff vs (B−V ) (Fig. 1), (B2−V1), and (B2−G) planes. In
fact, a large quadratic term (coefficient a5 in Table 2) for
these blue-optical colors may not be unphysical given the
large blanketing effects in the blue-visual region (Fig. 7).
The effect is intrinsic to the IRFM and not a numerical
artifact introduced by the calibrations.
4.5. Surface gravity effects
The effects of surface gravity (log g) on colors are illus-
trated in Fig. 8, where the difference between the color
of a dwarf and a giant star of the same Teff is plotted
against Teff for the (B − V ), (b − y), (V − J2), (V − H2),
and (V − K2) colors. A similar comparison for other
colors can be found in Fig. 17 of RM04a.
For Teff > 6000 K, decreasing log g has the effect of
increasing the absolute value of the slope of the Paschen
continuum (Fig. 7, see Sect. 5.2 in RM04a for a physi-
cal explanation). Since (B − V ) and (b − y) essentially
measure this (negative) slope, their values for giants will
be smaller (the slope becomes even more negative). This
is consistent with the differences plotted in Fig. 8 for
(B − V ) and (b − y).
Note that in the range 4800 K< Teff < 6000 K the
Johnson-2MASS colors are almost insensitive to log g
(the effect is less than 0.05 mag), which makes them suit-
able for stars of unknown luminosity class, in addition to
the fact that they are also nearly metallicity indepen-
dent. The influence of gravity on colors at low temper-
atures is very complex and still difficult to understand
(see Sect. 5.2 in RM04a).
4.6. Giants in clusters
In Part I, we derived the IRFM temperatures of a num-
ber of giants in the following clusters: M3, M67, M71,
M92, 47 Tuc, NGC 1261, NGC 288, and NGC 362. When
the calibrations for giants were performed, a preliminary
field calibration was first constructed, and then only the
clusters for which the internal cluster Teff scale was con-
sistent with the field Teff scale were included. This pro-
16
Ram´ırez & Mel´endez
Fig. 8. -- Color of a dwarf minus that of a giant at a fixed temperature as a function of Teff for: (B − V ) (solid line), (b − y) (dotted
line), (V − J2) (dashed line), (V − H2) (long dashed line), and (V − K2) (dot-dashed line).
cedure not only reduced the risk of systematic errors in-
troduced by errors in the metallicity and reddening cor-
rections of the clusters, but mainly avoided errors in the
stellar photometry. It is not surprising, then, that the
Teff scale for the DDO colors includes all the clusters
(see Fig. 18 in RM04a). The photometry is highly accu-
rate, and the metallicity effects on DDO colors strong.
This result confirms that the Teff scale of giants in the
field and that of giants in clusters is the same, and that
the metallicity and reddening scales of the clusters were
well determined.
However, there are still small discrepancies in the
metallicities of globular clusters, e.g.
for M71 Ram´ırez
et al. (2001) give [Fe/H] = −0.77, Sneden et al. (1994)
[Fe/H] = −0.79, and Carreta & Gratton (1997, here-
after CG97) [Fe/H] = −0.70; for NCG 362 Shetrone
& Keane (2000) suggest [Fe/H] = −1.33 and CG97
[Fe/H] = −1.15; while for M92 Sneden et al. (2000) ob-
tain [Fe/H] = −2.37, King et al. (1998) [Fe/H] = −2.52,
and CG97 [Fe/H] = −2.16. The values we adopted are
mainly those given by Kraft & Ivans (2003), which are
in reasonable agreement with the mean of the different
values cited in the literature.
The Teff vs. (B − V ) calibration does not include M3,
NGC 288, NGC 1261, and 47 Tuc due to large photomet-
ric uncertainties. The photometry for the other clusters
reproduce the (B − V ) Teff scale very well (Fig. 9). Even
for the RGB tip of M3 we found good agreement when
the best available photometry of three bright stars was
used, but the very large photometric errors of the fainter
giants produced a large disagreement between the two
scales.
5. COMPARISON WITH OTHER STUDIES
A considerable body of work exists on the Teff scale of
F, G, and K stars; some of these works have been dis-
cussed in Part I (Sect. 5.3). The IRFM Teff scale has
been compared with the results of several groups and
has been found, in general, in reasonable agreement with
them (AAM, MR03, RM04a). The reader is referred to
AAM, MR03, and RM04a to see how the present IRFM
Teff scale specifically compares to other results. Here we
will proceed to show that the present work is an update
of AAM work and our earlier extensions. We also revisit
the comparisons with the synthetic colors derived from
Kurucz models (M. S. Bessell 2004, private communica-
tion) and MARCS models (Houdashelt et al. 2000).
In Figs. 10 and 11 we compare our Teff vs. (B−V ) rela-
tion with that given by AAM for dwarfs of [Fe/H] = +0.0
Fig. 9. -- Teff vs.
(B − V ) relations for giants in the open
cluster M67 (filled circles) and in the globular clusters M71 (open
circles), NGC 362 (squares), and M92 (triangles). Solid, dotted,
dashed, and long-dashed lines correspond to our calibrations for
[Fe/H] = −0.08, [Fe/H] = −0.80, [Fe/H] = −1.30, and [Fe/H] =
−2.40, respectively, which are approximately the mean metallicities
adopted for the clusters. Typical errors bars (0.05 mag and 60 K)
are shown to the upper right corner of the figure.
and [Fe/H] = −2.5, and giants of [Fe/H] = +0.0 and
[Fe/H] = −2.0, respectively. These particular values of
[Fe/H] were adopted to maximize the ranges in common.
The [Fe/H] difference is also large enough as to easily
distinguish the solar-metallicity Teff vs. color relation
from the metal-poor one. The theoretical calibrations of
Bessell and collaborators (see e.g. Bessell et al. 1998), as
well as that of Houdashelt et al. (2000), are also shown
(for the dwarf comparison, colors for log g = 4.5 were
adopted, for giants we used the colors for log g = 2.5).
Our Teff vs. (B − V ) relations are essentially the same
as those of AAM. There are, however, small departures
at low temperatures (< 4500 K for dwarfs, < 4000 K for
giants) for solar metallicity, and around Teff = 6000 K
for the metal-poor end.
In the latter case, for a fixed
temperature, the colors from our Teff scale are redder by
about 0.03 mag. Note that AAM relations for dwarfs
are almost parallel, i.e. the effect of [Fe/H] appears to
be color (or temperature) independent. The synthetic
The effective temperature scale of FGK stars. II.
17
IRFM, the relations of [Fe/H] = +0.0 and [Fe/H] = −2.0,
in the giant Teff vs. (B − V ) plane, get closer as Teff is re-
duced. This behaviour is not in agreement with the the-
oretical calculations, which suggest an almost constant
[Fe/H] effect (Fig. 11). The solar metallicity prediction
of Houdashelt et al.
is in excellent agreement with the
present results except at the cool end.
In Fig. 12 we compare the C(42−48) and (B2 −V1) col-
ors as a function of Teff and [Fe/H] for both dwarf and
giant stars, as obtained in this work, with those from
our earlier work (MR03 and RM04a). The main differ-
ence is the updated temperatures of the present work, of
course; MR03 and RM04a are based on AAM tempera-
tures. Also, no polynomial corrections were performed
in MR03. The comparison shows that the two Teff scales
differ by less than 0.05 mag in C(42 − 48) and 0.03 mag
in (B2 − V1). These differences are well within the photo-
metric errors, systematic errors in the previous calibra-
tions, and errors in the temperatures. The differences
do not exhibit any significant trend with either Teff or
[Fe/H]. The differences are larger for the DDO color
calibration, which is due to the lower number of stars
defining the Teff vs. C(42 − 48) relation as compared
to the Teff vs. (B2 − V1) relation. The calibration for
(B2 − V1) is obviously more robust, but even in the case
of C(42 − 48) the differences are not large.
A comparison of our Teff vs. (V − IC ) and (V − K) re-
lations with those given by Bessell and Houdashelt et al.
is illustrated in Fig. 13. Predictions of solar metallicity
and [Fe/H] = −2.0 are shown. The (V − K) colors of
Bessell are in the Bessell & Brett (1988) system, those
by Houdashelt et al. are in the Johnson-Glass system;
and ours are in the Johnson-2MASS system. Their ex-
act values may be somewhat different, depending on the
color, but the metallicity effects on the colors should be
essentially the same.
In any case, the Bessell & Brett
(V − K) colors, for example, only need to be shifted by
0.04 mag to be transformed into the Johnson-2MASS sys-
tem (Carpenter 2001). The comparison for the (V − K)
colors is thus still meaningful.
At solar metallicity the ∆Teff/∆color gradients are in
reasonable agreement with our dwarf calibrations above
4500 K and with the giant calibrations below 5500 K. The
effect of the metallicity is remarkably different between
the two theoretical results for the (V − IC ) color at high
temperatures: while Houdashelt et al.
colors tend to
be redder with a lower [Fe/H], Bessell suggests a bluer
color. The latter is in better agreement with our results,
but only differentially, as a shift of about 150 K may be
required to match the two Teff scales. Regarding (V −K),
the general trends are very similar at low temperatures,
although the IRFM suggests a very strong metallicity
effect that makes the cool dwarfs very blue. They are
bluer also according to the theoretical results but the
effect there is not very strong. For Teff > 6000 K both
sets of synthetic colors agree in that the metal-poor stars
are redder, a result discussed in §4.4. Note, however, that
the synthetic (V − K) colors are even more metallicity
dependent.
Kurucz and MARCS colors are thus in reasonable qual-
itative agreement with the IRFM, but the synthetic col-
ors may be failing to predict the detailed metallicity
dependence. Even the solar metallicity colors are not
very well reproduced by model atmospheres and syn-
Fig. 10. -- Teff vs. (B − V ) relation for main sequence stars of
[Fe/H] = +0.0 (solid lines and filled circles) and [Fe/H] = −2.5
(dotted lines and open circles) according to:
this work (thick
black lines), Alonso et al. (circles), Bessell (thin cyan lines), and
Houdashelt et al. (magenta lines).
Fig. 11. -- Teff vs. (B − V ) relation for giant stars of [Fe/H] =
+0.0 (solid lines and filled circles) and [Fe/H] = −2.0 (dotted lines
and open circles) according to: this work (thick black lines), Alonso
et al.
(circles), Bessell (thin cyan lines), and Houdashelt et al.
(magenta lines).
colors obtained from both Kurucz and MARCS models,
however, show a quite complicated behaviour, the rela-
tions for solar metallicity and [Fe/H] = −2.5 are closer
both at the cool and hot extremes. The trends predicted
by model atmospheres at Teff > 6000 K are consistent
with our results. On the other hand, the synthetic col-
ors seem to be failing to reproduce the strong metallicity
effects for cool (Teff < 5000 K) dwarfs. According to
our IRFM Teff scale, at (B − V ) = 1.0 the tempera-
ture for [Fe/H] = 0.0 is about 400 K higher than that
at [Fe/H] − 2.5, but Houdashelt et al. suggest a ∆Teff of
only about +100 K.
AAM temperatures for giants are slightly cooler than
ours for Teff & 6500 K but they are in an almost perfect
agreement with ours everywhere else. According to the
18
Ram´ırez & Mel´endez
Fig. 12. -- Color difference at a given temperature between the Teff scale derived in this work and that derived in MR03 and RM04a,
which is based on Alonso et al. temperatures. Solid, dash-dotted, and dotted lines correspond to [Fe/H] = +0.0, [Fe/H] = −1.0, and
[Fe/H] = −2.0 main sequence stars, respectively. Filled circles, open circles, and triangles correspond to [Fe/H] = +0.0, [Fe/H] = −1.0,
and [Fe/H] = −2.0 giants, respectively.
thetic colors, since the predicted colors require not only
a zero point correction, but a correction that depends
also on spectral type (see e.g. Sect. 3.2.2 in Houdashelt
et al. 2000). In particular, the calibration of Houdashelt
et al., while in excellent agreement with the IRFM for so-
lar metallicity stars, shows differences as large as ±200 K
for stars with [Fe/H] = −2, but note that it depends on
the color being compared, metallicity, evolutionary stage
and spectral type. For example, for an F metal-poor
([Fe/H] = −2) dwarf, the agreement between the IRFM
and Houdashelt et al. (V −K) colors is very good, but for
cooler stars the temperatures derived from Houdashelt
et al. are systematically higher, and at (V − K) = 2.5,
Houdashelt et al.
temperatures are higher than those
from the IRFM by 200 K. These differences become es-
pecially important for samples covering a large range in
Teff; for example, when studying small abundance varia-
tions from the turn-off to the RGB tip of globular clus-
ters, spurious variations could be found for stars of dif-
ferent evolutionary stages. Fortunately, the problem is
alleviated when abundances with respect to iron are re-
ported ([X/Fe I]), unless the lines of the element X have
a temperature dependence very different from those of
Fe I.
6. CONCLUSIONS
Based on a large number of main sequence and gi-
ant stars with temperatures determined with the infrared
flux method (IRFM), a set of homogeneously calibrated
temperature vs. color relations is provided. The calibra-
tions include the effect of [Fe/H] and residual corrections
to avoid systematic effects introduced by the calibration
formulae, with the aim of reproducing the effects of spec-
tral features that cannot be taken into account by the
initial fits.
The calibrations have been tested with a sample of
stars with known direct and IRFM temperatures. Usu-
ally, more than eight colors were adopted to derive a pho-
tometric temperature. The comparison of photometric
temperatures with those from the IRFM shows excellent
agreement, with a dispersion fully consistent with the er-
rors in the IRFM temperatures. Thus, we have shown
that the calibrations do not introduce any systematic
error. When compared to the direct temperatures, not
only is good agreement found, but we have also shown
that the zero point of the IRFM temperature scale is
in agreement with the absolute zero point, to a level of
about 10 K.
The colors of the Sun, as determined from the calibra-
tions for a star of Teff = 5777 K and [Fe/H] = +0.0, are
presented and compared with those of five solar analogs.
A very good agreement is found considering the photo-
metric uncertainties, which makes these colors useful to
the search of solar twins in various surveys, particularly
Hipparcos-Tycho and 2MASS.
Metallicity and surface gravity effects on the IRFM
Teff scale may be reasonably understood with the help of
theoretical spectra, although the situation for K stars is
still difficult to address. For stars with Teff & 6000 K,
the adoption of a solar metallicity temperature derived
from the (V − K) color for a metal-poor star is unaccept-
able, as systematic errors of the order of 200 K may be
introduced.
Provided that the photometry is accurate, good agree-
ment is found when comparing the temperature scales of
giants in the field and giants in clusters. Thus, the IRFM
Teff scale is the same for both field and cluster giants.
As expected, our Teff scale closely resembles the one
that is based on the Alonso et al.
temperatures (our
work is largely based on their study). The present re-
sults, however, have a better coverage of the atmospheric
parameter space and are more reliable: better and up-
The effective temperature scale of FGK stars. II.
19
Fig. 13. -- Teff vs. (V − IC ) and (V − K) relations for dwarf (upper pannel) and giant (lower pannel) stars of [Fe/H] = +0.0 (solid lines)
and [Fe/H] = −2.0 (dotted lines) according to: this work (thick black lines), Bessell (thin cyan lines), and Houdashelt et al. (magenta
lines).
dated input data was adopted, and the calibrations were
carefully performed to avoid the introduction of numeri-
cal sources of error. A number of problems of interest to
stellar evolution and the chemical evolution of the Galaxy
depend on the assumptions made in color vs. Teff rela-
tions. Our calibrations will permit these problems to be
tackled with greater confidence.
JM thanks partial support from NSF grant AST-
0205951 to JG Cohen. We thank M. S. Bessell for pro-
viding colors for a more complete set of Kurucz models,
I. Ivans for improving the English language and style of
the manuscript, and the anonymous referee for comments
and suggestions that helped to improve the paper. This
publication makes use of data products from 2MASS,
which is a joint project of the University of Massachusetts
and IPAC/Caltech, funded by NASA and the National
Science Foundation; and the SIMBAD database, oper-
ated at CDS, Strasbourg, France.
REFERENCES
Allende Prieto, C., Barklem, P. S., Lambert, D. L., & Cunha, K.
Blackwell, D. E., Petford, A. D., & Shallis, M. J. 1980, A&A, 82,
2004, A&A, 420, 183
249
Alonso, A., Arribas, S., & Mart´ınez-Roger, C. 1996, A&A, 313, 873
(AAM)
Bonifacio, P., & Molaro, P. 1997, MNRAS, 285, 847
Cayrel de Strobel, G., Soubiran, C., & Ralite, N. 2001, A&A, 373,
Alonso, A., Arribas, S., & Mart´ınez-Roger, C. 1999, A&AS, 140,
159
261 (AAM)
Bersier, D. 1996, A&A, 308, 514
Bessell, M. S. 1979, PASP, 91, 589
Bessell, M. S. 1986, PASP, 98, 1303
Bessell, M. S. 2001, PASP, 113, 66
Bessell, M. S., & Brett, J. M. 1988, PASP, 100, 1134
Bessell, M. S., Castelli, F., & Plez, B. 1998, A&A, 333, 231
Cameron, L. M. 1985, A&A, 146, 59
Carpenter, J. M. 2001, AJ, 121, 2851
Carretta, E., & Gratton, R. G. 1997, A&AS, 121, 95 (CG97)
Coc, A., Vangioni-Flam, E., Descouvemont, P., et al. 2004, ApJ,
600, 544
Cousins, A. W. J. 1976, Mem. R. Astron. Soc. 81, 25
Crawford, D. L. 1975, PASP, 87, 481
20
Ram´ırez & Mel´endez
Cutri, R. M., Skrutskie, M. F., Van Dyk, S., et al. 2003, VizieR
Mermilliod, J. C., Mermilliod, M., & Hauck, B. 1997, A&AS, 124,
Online Data Catalog, II/246
349 (GCPD)
Dean, J. F., Warren, P. R., & Cousins, A. W. J. 1978, MNRAS,
Mozurkewich, D., Armstrong, J. T., Hindsley, R. B., et al. 2003,
183, 569
Dawson, D. W. 1978, AJ, 83, 1424
ESA. 1997, The Hipparcos and Tycho Catalogues, ESA SP-1200
Fiorucci, M., & Munari, U. 2003, A&A, 401, 781
Francois, P., Depagne, E., Hill, V., et al. 2003, A&A, 403, 1105
Girardi, L., Bertelli G., Bressan, A., et al. 2002, A&A, 391, 195
Golay, M. 1966, in: Proceedings from the IAU Symposium no. 24,
Sweden, 1964, eds. K. Loden et al., Academic Press, London, p.
262
Gratton, R. G., Carretta, E., Claudi, R., et al. 2003, A&A, 404,
187
Hill, V., Plez, B., Cayrel, R., et al. 2002, A&A, 387, 560
Houdashelt, M. L., Bell, R. A., & Sweigart, A. V. 2000, AJ, 119,
1448
Johnson, H. L., & Morgan, W. W. 1953, ApJ, 117, 313
Kararas, G., Straizys, V., Sudzius, J., & Zdanavicius, K. 1966, Bull.
Vilnius Obs., No. 22, 3
King, J. R., Stephens, A., Boesgaard, A. M., & Deliyannis, C. 1998,
AJ, 115, 666
Kraft, R., & Ivans, I. 2003, PASP, 115, 143
Kuriliene, G., & Sudzius, J. 1974, Bull. Vilnius Obs., No. 40, 10
Lejeune, Th., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229
Magain, P. 1987, A&A, 181, 323
Mart´ınez-Roger, C., Arribas, S., & Alonso, A. 1992, Societa
Astronomica Italiana, Memorie, 63, 2, 263-281
McCall, M. L. 2004, AJ, 128, 2144
McClure, R. D., & van den Bergh, S. 1968, AJ, 73, 313
Mel´endez, J., Barbuy, B., Bica, E., et al. 2003, A&A, 411, 417
Mel´endez, J., & Ram´ırez, I. 2003, A&A, 398, 705 (MR03)
Mel´endez, J., & Ram´ırez, I. 2004, ApJ, 615, L33
AJ, 126, 2502
Porto de Mello, G. F., & da Silva, L. 1997, ApJ, 482, L89
Ram´ırez, S. V., Cohen, J. G., Buss, J., & Briley, M. M. 2001, AJ,
122, 1429
Ram´ırez, I., & Mel´endez, J. 2004a, A&A, 417, 301 (RM04a)
Ram´ırez, I., & Mel´endez, J. 2004b, ApJ, 609, 417
Reddy, B. E., Tomkin, J., Lambert, D. L., & Allende Prieto, C.
2003, MNRAS, 340, 304
Romano, D., Tosi, M., Matteucci, F., & Chiappini, C. 2003,
MNRAS, 346, 295
Ryan, S. G., Norris, J. E., & Beers, T. C. 1999, ApJ, 523, 654
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Sekiguchi, M., & Fukugita, M. 2000, AJ, 120, 1072
Shetrone, M. D. 1996, AJ, 112, 1517
Shetrone, M. D., & Keane, M. J. 2000, AJ, 119, 840
Sneden, C., Kraft, R. P., Langer, G. E., et al. 1994, AJ, 107, 1773
Sneden, C., Pilachowski, C. A., & Kraft, R. P. 2000, AJ, 120, 1351
Sneden, C., Cowan, J. J., Lawler, J. E., et al. 2003, ApJ, 591, 936
Soubiran, C., & Triaud, A. 2004, A&A, 418, 1089
Straizys, V. 1995, Multicolor Stellar Photometry, Pachart
Publishing House, Tucson, p. 372
Straizys, V., & Sviderskiene, Z. 1972, A&A, 17, 312
Stromgren, B. 1966, ARA&A, 4, 433
Sudzius, J., Bobinas, V., Raudeliunas, S. 1996, Baltic Astronomy,
5, 485
Takeda, Y., Ohkubo, M., & Sadakane, K. 2002, PASJ, 54, 451
Taylor, B. J. 1986, ApJS, 60, 577
Yi, S. K., Kim, Y., & Demarque, P. 2003, ApJS, 144, 259
|
astro-ph/0607242 | 1 | 0607 | 2006-07-11T21:43:11 | AEGIS: The Diversity of Bright Near-IR Selected Distant Red Galaxies | [
"astro-ph"
] | We use deep and wide near infrared (NIR) imaging from the Palomar telescope combined with DEEP2 spectroscopy and Hubble Space Telescope (HST) and Chandra Space Telescope imaging to investigate the nature of galaxies that are red in NIR colors. We locate these `distant red galaxies' (DRGs) through the color cut (J-K)_{vega} > 2.3 over 0.7 deg^{2}, where we find 1010 DRG candidates down to K_s = 20.5. We combine 95 high quality spectroscopic redshifts with photometric redshifts from BRIJK photometry to determine the redshift and stellar mass distributions for these systems, and morphological/structural and X-ray properties for 107 DRGs in the Extended Groth Strip. We find that many bright (J-K)_{vega}>2.3 galaxies with K_s<20.5 are at redshifts z<2, with 64% between 1<z<2. The stellar mass distributions for these galaxies is broad, ranging from 10^{9}-10^{12} M_solar, but with most z>2 systems massive with M_*>10^{11} M_solar. HST imaging shows that the structural properties and morphologies of DRGs are also diverse, with the majority elliptical/compact (57%), and the remainder edge-on spirals (7%), and peculiar galaxies (29%). The DRGs at z < 1.4 with high quality spectroscopic redshifts are generally compact, with small half light radii, and span a range in rest-frame optical properties. The spectral energy distributions for these objects differ from higher redshift DRGs: they are bluer by one magnitude in observed (I-J) color. A pure IR color selection of high redshift populations is not sufficient to identify unique populations, and other colors, or spectroscopic redshifts are needed to produce homogeneous samples. | astro-ph | astro-ph |
AEGIS: The Diversity of Bright Near-IR Selected Distant Red
Galaxies
C. J. Conselice1, J. A. Newman2,3, A. Georgakakis4, O. Almaini1, A. L. Coil3,5, M.C.
Cooper6, P. Eisenhardt7, S. Foucaud1, A. Koekemoer8, J. Lotz9, K. Noeske10, B. Weiner11,
C.N.A Willmer5
ABSTRACT
We use deep and wide near infrared (NIR) imaging from the Palomar telescope
combined with DEEP2 spectroscopy and Hubble Space Telescope (HST) and
Chandra Space Telescope imaging to investigate the nature of galaxies that are
red in NIR colors. We locate these 'distant red galaxies' (DRGs) through the
color cut (J − K)vega > 2.3 over 0.7 deg2, where we find 1010 DRG candidates
down to Ks = 20.5. We combine 95 high quality spectroscopic redshifts with
photometric redshifts from BRIJK photometry to determine the redshift and
stellar mass distributions for these systems, and morphological/structural and
X-ray properties for 107 DRGs in the Extended Groth Strip. We find that many
bright (J − K)vega > 2.3 galaxies with Ks < 20.5 are at redshifts z < 2,
with 64% between 1 < z < 2. The stellar mass distributions for these galaxies
is broad, ranging from 109 − 1012 M⊙ , but with most z > 2 systems massive
with M∗ > 1011 M⊙ . HST imaging shows that the structural properties and
morphologies of DRGs are also diverse, with the majority elliptical/compact
1School of Physics and Astronomy, University of Nottingham, NG7 2RD, UK
2Lawrence Berkeley National Laboratory, Berkeley, CA 94720
3Hubble Fellow
4Imperial College, London
5Stewart Observatory, University of Arizona, Tucson, AZ
6University of California, Berkeley
7NASA Jet Propulsion Laboratory, California Institute of Technology
8Space Telescope Science Institute, Baltimore, MD
9Goldberg Fellow, National Optical Astronomy Observatory, Tucson, AZ 85726
10University of California, Santa Cruz, Santa Cruz, CA
11Department of Physics, University of Maryland, College Park, MD 20742
– 2 –
(57%), and the remainder edge-on spirals (7%), and peculiar galaxies (29%). The
DRGs at z < 1.4 with high quality spectroscopic redshifts are generally compact,
with small half-light radii, and span a range in rest-frame optical properties.
The spectral energy distributions for these objects differ from higher redshift
DRGs: they are bluer by one magnitude in observed (I − J) color. A pure IR
color selection of high redshift populations is not sufficient to identify unique
populations, and other colors, or spectroscopic redshifts are needed to produce
homogeneous samples.
1.
Introduction
Uncovering the formation mechanisms of the most massive galaxies has a long history
and has largely driven the field of galaxy formation and evolution. Traditional analyses
of stellar populations in nearby ellipticals reveal that the bulk of the stars in the most
massive galaxies formed early, within a few Gyr of the Big Bang (e.g., Trager et al. 2000).
Observing distant galaxies is complementary to nearby galaxy age-dating, as it allows us to
directly observe galaxies while they are forming. Early detections of high redshift galaxies
revealed that UV bright star forming galaxies at z ∼ 3 are dominated by relatively low
stellar mass systems (∼ 1010 M⊙ ) (Papovich et al. 2005). As the most massive galaxies
today are > 1011.5 M⊙ , this implies that high−z galaxies grow by a factor of 5-10 through
some process, such as merging (e.g., Conselice et al. 2003; Conselice 2006).
It seems likely that at least some progenitors of today's massive galaxies are not bright
in the UV, and a population of near infrared (NIR) selected 'Distant Red Galaxies' (DRGs)
was described by Saracco et al. (2001) and Franx et al. (2003), who suggested that some
of these systems are the progenitors of the most massive present day galaxies. These DRGs
are selected using a single color cut, (J − K)vega > 2.3, and thus require near infrared (NIR)
imaging to detect. While it has been proposed that DRGs are massive z > 2 galaxies, the
bright end of this population has yet to be studied, and thus we do not yet know the general
characteristics of red galaxies selected in the near infrared.
We investigate this problem by examining the redshifts, masses, morphologies and spec-
tra of DRGs in the Extended Groth Strip (EGS). We study 1010 bright DRGs with Ks < 20.5
within the EGS, and two other fields that are part of our NIR survey with the Palomar tele-
scope. We examine the structures and morphologies of DRGs that are coincident between
the ACS imaging in the EGS and our NIR survey. We find that a high fraction of DRGs at
Ks < 20.5 are at z < 2, with the majority between 1 < z < 2, including 95 (10%) with spec-
troscopic redshifts. We present an initial study of these NIR selected galaxies at z < 1.4 and
– 3 –
conclude that they are small, low mass, AGN/star formation dominated galaxies. Further-
more, we find a diversity in the morphological properties of DRGs at all redshifts, suggesting
that there is a wide diversity of properties for galaxies selected with the (J − K)vega > 2.3
color cut, perhaps even larger than that seen for traditional EROs (Moustakas et al. 2004).
In this paper we assume the following cosmology: H0 = 70 km s−1 Mpc−1, Ωλ = 0.7, and
Ωm = 0.3, and use Vega magnitude units throughout.
2. Data and Sample
The data used in this paper originate from multiple data sets partially associated with
the All-wavelength Extended Groth Strip International Survey (AEGIS) (Davis et al. 2006).
The main source of data is from the Palomar Observatory Wide-Field Infrared Survey
(POWIR) (Bundy et al. 2005a,b; Conselice et al.
in prep), while other sources include
DEEP2 spectroscopy (Davis et al. 2003), one orbit Hubble Space Telescope (HST) imaging
using the Advanced Camera for Surveys (ACS) of the EGS (Lotz et al. 2006), and an X-ray
survey of the region (Nandra et al. 2006 in prep). The POWIR data were obtained from
September 2002 until October 2005 at the Palomar Observatory 5 meter using the Wide-field
Infrared Camera (WIRC). For the J and K data we cover a total of ∼ 0.7 deg2 to 5-σ point
source sensitivities of Ks = 20.5 − 21.5 and J = 22.5 − 23. Within the EGS, where our
highest quality data exist, we image 0.2 deg2 to a 5 sigma depth of Ks = 21.1 and J = 23
for point sources. To construct a sample of DRGs which is nearly complete in the K-band,
we limit our study of DRGs to those at Ks < 20.5.
We select DRGs through the use of a (J − K)vega > 2.3 cut. To supplement the existing
95 DEEP2 spectroscopic redshifts, we calculate photometric redshifts derived from BRIJK
photometry using the ANNz method for z < 1.4 systems, and hyper-z for those at z > 1.4.
The photometric redshift accuracy at z < 1.4 for all galaxies is quite good, with δz/z =
0.05, based on comparisons to > 10, 000 spectroscopic redshifts. The subset of 95 DRGs
with z < 1.4 spectroscopy have a slightly softer accuracy of δz/z = 0.07, with very few
catastrophic mismatches. We test our z > 1.4 photometric redshifts by comparing to 44
galaxies in the Steidel et al. (2004) BM/BX 1.5 < z < 2.5 population, where we find a lower
accuracy of δz/z = 0.22.
We use the spectral energy distributions (SEDs) and spectroscopic/photometric red-
shifts to calculate stellar masses utilizing the methods outlined in Bundy et al. (2005a,b).
Our star formation histories are parameterized by a single exponentially declining star for-
mation event. We vary the metallicity, dust extinction and age of these various models to
calculate the distribution of stellar masses using a Chabrier IMF (see Bundy et al. 2005a,b
for a detailed discussion).
– 4 –
3. Results
3.1. Basic Redshift and Mass Characteristics
We find a total of 1010 DRGs, brighter than Ks = 20.5, over our 0.7 deg2 (2520 arcmin2)
Palomar survey area, including two fields in addition to the EGS. In comparison to previous
work, van Dokkum et al. (2006) studied ∼ 200 DRGs over 400 arcmin2, FIRES 14 DRGs
over 100 arcmin2, and Papovich et al. (2006) 153 DRGs over 130 arcmin2, all with similar
DRG number densities at Ks < 20.5 as found in our study. Out of our 1010 NIR selected
DRGs, we determine that 55 are stars based on their optical colors and unresolved structures.
The DRG redshift distribution is shown in Figure 1. A total of 95 of these DRGs have
secure spectroscopic redshifts taken with the DEIMOS spectrograph as part of the DEEP2
survey (Davis et al. 2003). This gives us a lower limit of ∼ 10% on z < 1.4 galaxies in
pure photometrically selected DRG samples to Ks = 20.5. Using these redshifts with our
calibrated photometric redshifts at z < 1.4, we find that ∼ 70% of Ks < 20.5 DRGs are at
z < 1.4, but most of the DRG population (64%) is between 1 < z < 2. Objects at z > 2
represent ∼ 4% of the Ks < 20.5 DRG population, while ∼ 25% of the population is found
between 0.5 < z < 1. This implies that bright DRGs have a broad redshift distribution,
which is also found in other recent studies (Reddy et al. 2005; Papovich et al. 2006; Grazian
et al. 2006).
We plot the stellar mass distribution for our sample in Figure 1b. From this we conclude
that (J − K)vega > 2.3 selected galaxies do not uniquely sample high-mass galaxies, with a
low-redshift 'contamination' rate approximately similar to what Franx et al. (2003) estimate
based on Smail et al. (2002). DRGs however do select high redshift systems, generally at
z > 0.8, and have a redshift distribution broader than the ERO population (Moustakas et al.
2004). A (J − K)vega > 2.3 selection at z ∼ 1.5 measures approximate (B − I) colors similar
to dusty or old galaxies at those redshifts, based on 6 Gyr single burst stellar population
models. In Figure 2 we plot the (J − K) vs. Ks diagram for all DRGs at z > 1.5 that have
stellar masses > 1011 M⊙ . The most massive galaxies with masses > 1011.5 M⊙ are shown as
red boxes for 1.5 < z < 2.0 systems, and blue circles for those at z > 2. The green crosses
represent galaxies with M∗ > 1011 M⊙ at z > 2. The average (J − K) color for all galaxies
at z > 2 with M∗ > 1011 M⊙ is < (J − K) >= 2.3 ± 0.62, while the most massive systems
with M∗ > 1011.5 M⊙ have a mean color of < (J − K) >= 2.4 ± 0.7. Although the average
massive galaxy has a similar color as the DRG limit, we find that a significant number
(∼ 50%) of all massive galaxies at 2 < z < 3 will be missed by the standard DRG selection,
– 5 –
in agreement with van Dokkum et al. (2006). These results are robust when using brighter
magnitude cuts where our photometry is more accurate, and at redder color limits, up to
(J − K)vega > 3.3. These trends also remain after we select only galaxies that are redder
than 1 σ of the DRG limit, and after testing our selection through a bootstrap resampling.
That is, the diversity in redshifts/masses for bright DRGs is not produced by photometric
errors bringing up galaxies into the DRG cut.
We find however that at z > 2 the DRG selection limit to Ks = 20.5 does locate very
massive galaxies, with an average stellar mass of <M∗> = 1011.7±0.2 M⊙ . The average DRG
at all redshifts has a stellar mass of M∗ = 1010.8±0.5 M⊙ , with the spectroscopic sample at
z < 1.4 having an average stellar mass of M∗ = 1010.5±0.3 M⊙ . For the remainder of this
paper we focus on the DRGs found within the EGS region, where we have ancillory HST
and X-ray data.
3.2. Structures and Morphologies of DRGs
Due to the ACS overlap with our NIR imaging, we can determine the morphological
properties of 107 DRGs at Ks < 20.5 up to z ∼ 4. We utilize both apparent eye-ball mor-
phologies and quantitative measurements of size and structure on the F814W ACS imaging
in the EGS to determine what type of galaxies DRGs are. We classify by eye our galaxies
into the types (with the number of objects in each class): ellipticals/compact (61), peculiars
(31), disks (0), edge-on disks (7), and those too faint to classify (8) following the method
described in Conselice et al. (2005). The morphologies for the (J − K)vega > 2.3 galaxies
show a diversity of types, as can be seen in Figure 3, similar to the situation for lower redshift
EROs (Moustakas et al. 2004). The most common type are ellipticals/compact making up
57% of the total DRG sample, with most of these (80%) compact. About 7% of the sam-
ple are composed of edge-on disks, which are likely red in (J − K) due to dust redenning.
Peculiars account for the remainder (29%) of the DRG sample. At z > 2 there is roughly
a similar mixture of ellilptical and peculiars. We have morphologies for 14 systems with
spectroscopic redshifts, the majority of which are compact (§4). We find very little evolution
in the morphological distribution with redshift, with a slight increase in the relative number
of peculiar galaxies at higher redshift.
We also utilized the revised CAS system (Conselice 2003; Conselice et al. 2004) to
quantify the structures of the (J − K)vega > 2.3 galaxies. The CAS (concentration, asym-
metry, clumpiness) parameters are a non-parameteric method for measuring the structures
of galaxies as resolved on CCD images (Conselice 2003). As expected from the eye-ball clas-
sifications, we find a diversity of CAS values. For the systems with spectroscopic redshifts at
– 6 –
z < 1.4 we find average values, < C >= 2.4 ± 0.1, < A >= 0.26 ± 0.11, < S >= 0.09 ± 0.06,
which is typical for nearby normal galaxies, and spirals with exponential profiles (Conselice
2003). We find nearly the same values for galaxies with z < 1.4 photometric redshifts.
For the higher redshift sample at z > 1.5 we find that the morphologically classified el-
lipticals/compacts are more concentrated with < C >= 3.0 ± 0.5, with many systems at
C > 3.3, similar in morphology to nearby massive ellipticals. Furthermore, as we are exam-
ining z > 1.4 systems in the UV, we are likely underestimating the rest-frame optical light
concentration (Taylor et al. 2006, submitted). This is an indication, along with their high
masses, that these z > 1.5 systems are likely elliptical progenitors.
There are two caveats to using the F814W band ACS imaging on these galaxies. The
first is that there are redshift effects which will change the measured parameters (Conselice
et al. 2000a,b; Conselice 2003). Systems at z > 1.2 are also viewed in the rest-frame
ultraviolet, complicating comparisons to nearby galaxies viewed in the optical. There is
evidence, however, that distant galaxies dominated by UV bright star formation look similar
in the rest-frame optical and UV (Windhorst et al. 2002; Papovich et al. 2005; Conselice et
al. 2005).
4. Low Redshift z < 1.4 DRGs
We have 95 confirmed high quality spectroscopic redshifts for DRGs at z < 1.4. About
7% of these systems are traditional EROs with (R − K)vega > 5, while ∼ 33% have red
optical/NIR colors, with (R − K)vega > 4. The average magnitude of our spectroscopically
confirmed lower redshift DRGs at z < 1.4 is <MB >= −20.48±1.27, with an average color of
< (U −B) >= −0.16±0.23, and these systems span the color-magnitude relation, suggesting
a heterogeneous origin. Our lower redshift DRGs are a mangitude bluer in observed (I − J)
color compared to the z > 1.5 DRGs. They are also roughly a magnitude brighter than the
DEEP2 spectroscopic limit, with an average <R> = 23.3. The average half-light radii of the
spectroscopically selected (J − K)vega > 2.3 galaxies is 0.7±1 kpc, with an average stellar
mass of 109.8±0.46 M⊙ within the ACS region. The concentration indices of these systems are
moderate, with a average of C = 2.8±0.4, typical for low mass ellipticals, or disks (Conselice
2003). The systems at z < 1.4 with photometric redshifts in the ACS region have very
similar masses, sizes and morphologies as the spectroscopic sample. Interestingly, none of
these lower redshift DRGs are face-on disks, which is the most common galaxy type at
z < 1.4 (Conselice et al. 2005).
To further investigate the nature of the z < 1.4 objects we combine the spectra of
all DRGs with spectroscopy to produce a co-added non-flux weighted (Figure 4) spectrum.
– 7 –
This spectrum shows that the z < 1.4 DRGs with emission lines, host star formation, AGN
activity, as well as evidence for post starbursts with ages 1 Gyr, revealed through strong
Balmer absorption lines. Features produced by star formation, including [OII] and Balmer
absorption lines, are seen, as well as higher excitation lines such as [NeIII] and [OIII]. The
average ratio of ([OIII]λ5007)/(Hβ λ4861) ∼ 4 −4.5 after correcting for Hβ absorption. This
suggests a mixture of star formation and AGNs (Seyfert 2s) could be responsible for the lower
redshift sources (Veilleux & Osterbrock 1987). The preliminary X-ray Chandra catalog of
the EGS reveals that a lower limit of six z < 2 DRGs have bright X-ray detections to a limit
of ∼ 10−16 erg s−1 cm−2, all with photo-zs, and soft hardness ratios. Other sources could
be obscured AGN in moderate Lx systems. We conclude that (J − K)vega > 2.3 systems at
z < 1.4 appear to have average masses, a moderate concentration of light, and are host to
star formation and AGN activity, but otherwise are heterogenous.
5. Discussion
With several large area and deep NIR surveys coming online soon, such as UKIDSS
and VISTA, it is desirable to understand the properties of pure photometrically selected
NIR samples, such as the DRGs with (J − K)vega > 2.3. Our results suggest that bright
DRGs are a mixed population, as other smaller studies have previously found (Reddy et al.
2005; Papovich et al. 2006; van Dokkum et al. 2006). The differences between the original
DRG, Franx et al. (2003), study and ours is due to the depth of the surveys, with Franx
et al. being several magnitudes deeper, although most DRGs studied in detail thus far are
at Ks < 20. Because DRGs have unique SEDs, their photometric redshifts are potentially
harder to measure accurately, and these results, and others that utilize photometric redshifts,
should be viewed as tentative until significant numbers of spectroscopic redshifts become
available. This is demonstrated by a high absolute lower limit (10%) contribution to the
entire DRG population from z < 1.4 galaxies, which tend to be small, lower mass galaxies
with optical AGN signatures. The morphological mix derived in this paper is however robust,
regardless of the redshift distribution. The DRG population is morphologically dominated
by compact galaxies, with edge-on spirals and peculiars making up 36% of the population. In
the future multi-object near-IR spectrographs will be necessary to make definitive progress
in our understanding of non-UV bright galaxies at z > 2.
We thank the members of the AEGIS, Palomar and DEEP2 teams, particularly Kevin
Bundy, for their many invaluable contributions to the surveys that have made this paper pos-
sible. We thank Chuck Steidel for allowing us to utilize the spectroscopic redshifts from his
BX/BM survey before publication, and Casey Papovich and Naveen Reddy for enlightening
– 8 –
discussions. This work was supported by a NSF Astronomy & Astrophysics Postdoctoral
Fellowship, PPARC and NASA. ALC is supported by NASA through Hubble Fellowship
grant HF-01182.01-A.
REFERENCES
Bundy, K., Ellis, R.S., & Conselice, C.J. 2005a, ApJ, 625, 621
Bundy, K., et al. 2005b, astro-ph/0512465
Conselice, C.J., Gallagher, J.S., Calzetti, D., Homeier, N., & Kinney, A. 2000a, AJ, 119, 79
Conselice, C.J., Bershady, M.A., & Jangren, A. 2000b, ApJ, 529, 886
Conselice, C.J. 2003, ApJS, 147, 1
Conselice, C.J., Bershady, M.A., Dickinson, M., & Papovich, C. 2003, AJ, 126, 1183
Conselice, C.J., et al. 2004, ApJ, 600, 139L
Conselice, C.J., Blackburne, J., & Papovich, C. 2005, ApJ, 620, 564
Conselice, C.J. 2006, ApJ, 638, 686
Davis, M., et al. 2003, SPIE, 4834, 161
Franx, M. et al. 2003, ApJ, 587, 79L
Grazian et al. 2006, astro-ph/0603095
Lotz, J., et al. 2006, astro-ph/0602088
Moustakas, L.A., et al. 2004, ApJ, 600, 131L
Papovich, C., Dickinson, M., Giavalisco, M., Conselice, C.J., & Ferguson, H.C. 2005, ApJ,
631, 101
Papovich, C. et al. 2006, ApJ, 640, 92
Reddy, N.A., Erb, D.K., Steidel, C.C., Shapley, A.E., Adelberger, K., & Pettini, M. 2005,
ApJ, 633, 748
Saracco, P. et al. 2001, A&A, 375, 1
Smail, I., Owen, F.N., Morrison, G.E., Keel, W.C., Ivison, R.J., & Ledlow, M.J. 2002, ApJ,
581, 844
Trager, S.C., Faber, S.M., Worthey, G., & Gonzalez, J.J. 2000, AJ, 119, 1645
van Dokkum, P.G., et al. 2006, ApJ,
Veilleux, S., & Osterbrock, D.E. 1987, ApJS, 63, 295
Windhorst, R., et al. 2002, ApJS, 143, 113
– 9 –
This preprint was prepared with the AAS LATEX macros v5.2.
– 10 –
Fig. 1.- Redshift and stellar mass distributions for the Ks < 20.5, DRG (J − K)vega > 2.3
population over 0.7 deg2. The shaded red histogram in the redshift (left) panel is for galaxies
with reliable spectroscopic redshifts. The solid shaded red histogram for the mass (right)
panel is for galaxies with spectroscopic redshifts, while the blue dashed shaded histogram
shows the stellar mass distribution for galaxies at z > 2.
– 11 –
Fig. 2.- The (J − K)vega vs. Ks diagram for objects in the total Palomar coverage areas.
Every point is for an object at z > 1.5 which has a measured stellar mass with M∗ > 1011
M⊙ . Objects surrounded by red boxes are objects with M∗ > 1011.5 M⊙ at 1.5 < z < 2.0,
while circled objects are systems with M∗ > 1011.5 M⊙ at z > 2.0. Objects at z > 2 and with
M∗ > 1011 M⊙ are plotted as green crosses. The horizontal line is the limit for DRGs. The
typical error is shown in the upper left.
– 12 –
E/Compact
Ell
Pec
Pec.
Edge-On
Edge−on Disk
Fig. 3.- The morphological type fractions for DRGs with Ks < 20.5 in the EGS. The
fractions are displayed as: a solid green line for the compact/ellipticals, a long-dashed line
for peculiars, and a short-dashed red line for edge-on disks. Examples of each morphological
type are shown on the right hand side.
– 13 –
Fig. 4.- The co-added spectrum for the lower redshift DRGs with a high quality spectra
from the DEEP2 survey. Lines due to star formation, such as Hβ can be seen in tandem
with high ionization lines, such as [NeIII], which suggests the presence of AGN activity. The
ratio of ([OIII]λ5007)/(Hβ λ4861) is also high, which is also indicative of AGN excitation
for some systems.
|
astro-ph/9702003 | 1 | 9702 | 1997-01-31T06:41:03 | Multipole Expansion Model in Gravitational Lensing | [
"astro-ph"
] | Non-transparent models of multipole expansion model and two point-mass model are analyzed from the catastrophe theory. Singularity behaviours of $2^n$-pole moments are discussed. We apply these models to triple quasar PG1115+080 and compare with the typical transparent model, softened power law spheroids. Multipole expansion model gives the best fit among them. | astro-ph | astro-ph | Multipole Expansion Model in Gravitational
Lensing
Takeshi FUKUYAMA ∗, Yuuko KAKIGI ∗, and Takashi OKAMURA †
∗ Department of Physics,
Ritsumeikan University, Kusatsu,
Shiga, 525 Japan
† Department of Physics,
Tokyo Institute of Technology,
Oh-Okayama, Meguro-ku, Tokyo, 152 Japan
Abstract
Non-transparent models of multipole expansion model and two point-
mass model are analyzed from the catastrophe theory. Singularity be-
haviours of 2n-pole moments are discussed. We apply these models to triple
quasar PG1115+080 and compare with the typical transparent model, soft-
ened power law spheroids. Multipole expansion model gives the best fit
among them.
Key words : gravitational lensing, multipole expansion model.
7
9
9
1
n
a
J
1
3
1
v
3
0
0
2
0
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
1
Introduction
Gravitational lensing is and will be playing very important roles in astrophysics
[1] [2]. It covers the large region in scale from the brown dwarf to the supercluster
of galaxies. Microlensing uncovers the dark matter in the halo and the planetary
system in the neighbourhood of the Sun [3]. In this letter we deal with a special
macrolensing, that is, the multiple quasar supposed to be caused by a single
galaxy in the foreground. It enables us to reveal the structure of the lensing galaxy
without the ambiguity of surrounding galaxies. It also gives us the information of
the large scale structure of the universe through the angular diameter distance.
One of the most popular models of lensing galaxy is the spheroid with softened
power law behaviours [4]. The physical implication of this model is very clear
and it is easy to compare with observations, such as luminosity distribution and
ellipticity, etc. However these merits are also the demerits of this model since it
is too restrictive to model shape. We are not sure whether the lensing galaxy is
elliptical or spiral. It may be even complex system composed of several galaxies.
This model is also analytically non-tractable. On the other hand multipole ex-
pansion model (hereafter MPE model) is more general than the spheroid model.
MPE model is applicable to any complex system as far as the lensing objects are
compact with respect to impact parameters. Also it is analytically tractable.
Using it, we are able to estimate the structure of the lensing object systemat-
ically. That is, the more we incorporate multipole components, the more we can
know the detailed structure of the lensing object.
From the observational side, gravitational lensing effect seems to be very
promissing. Deep space survey by HST is now working and Sloan Digitized Sky
Survey will work soon [5]. However the present status of observation is not still
sufficient to uncover the affirmative results. For instance, parameter fitting is
very sensitive to the image positions [6] whose observation cannot be precise in
many cases. We need more detailed and more precise data to be adjusted. In
such a situation it is also very useful to study the same lensing phenomena by
several models, which reveal the common and therefore true features of lensing
objects.
In the previous paper [7] we discussed about the relationship between multi-
pole moments and caustic singularities, and applied it to PG1115+080. In this
treatise it is also shown that complex coordinates are very useful.
The merit of MPE model is that we can probe into the detailed structure
systematically by steping up higher order of multipoles. An arbitrarily complex
but compact lensing object is analytically tractable as far as each 2n-pole mo-
ment is discussed separately. So in this article we study the structure of 2n-pole
term systematically. If we sum up all contributions, its treatment goes beyond
analytical survey and is devoted to the numerical calculations.
Two point-mass lens is situated at the antipodal point of multipole expansion
model.
2
Two point-mass model consists of a series of multipole moments. However
it allows analytical survey as a whole though a straightforward application of
catastrophe theory still needs numerical calculations. In order to understand the
essential physical implication of any event, analytical survey is indispensable.
We apply the above mentioned models to PG1115+080 and discuss their
common and discrepant characters among them.
This paper is organized as follows. In section two we review the metamor-
phoses of caustic for later use whose detailed arguments are given in our previous
work [7]. This metamorphoses are applied to MPE model in section three and to
the two point-mass model in section five. Section four is devoted to the analysis
of the detailed structure of 2n-pole contribution in MPE model. MPE model is
applied to the multiple lens system of PG1115+080 in section six. The obtained
numerical results are compared with other models, especially with elliptical lens
model in section seven. Section eight is devoted to discussions.
2 Metamorphoses of the Caustics
For later studies, we summarize the metamorphoses of the caustics. This is
mainly based on our previous work [7].
Using complex variables, singularities are classified in compact form.
Critical line is
D ≡ −4det(φαβ)
= 4(φz ¯z2 − φzz2) = 0 ,
(α, β = z, ¯z)
(2.1)
where φ is Fermat potential and z is complex representation of image position,
z = x1 + ix2. ¯z is complex conjugate of z. Subscripts α, β on φ denotes the
derivative ∂α∂βφ and so on.
Here we use the Poisson equation
φz ¯z =
1
2
(1 − κ) .
κ is the normalized surface mass density and is equal to zero on the light path
for non-transparent model which we will consider in this article.
Then the critical line is reduced to
φzz ±
1
2
= 0
(2.2)
(φ22 = 0).
by the diagonalization in real coordinates. + (−) corresponds to φ11 = 0
Cuspoid sequences are characterized by
3
L is the derivative operator along the critical line, which is given by
Lnφz = 0
(n = 1, 2, 3) .
Their explicit forms are
L =
i
8
(∂¯zD∂z − ∂zD∂¯z) .
∂3φ
∂z3 ±
∂3φ
∂ ¯z3 = 0
for the cusp (n = 1) singularity,
∂4φ
∂z4 +
for the swallowtail (n = 2) and
(
∂4φ
∂ ¯z4 ) ± 6(
∂3φ
∂z3 )2 = 0
(
∂5φ
∂z5 ±
∂5φ
∂ ¯z5 ) ± 20
∂3φ
∂z3 {
∂4φ
∂z4 ± 3(
∂3φ
∂z3 )2} = 0
(2.3)
(2.4)
(2.5)
(2.6)
(2.7)
for the butterfly (n = 3).
It should be noticed that the higher cuspoids must also satisfy the conditions
of the lower ones. For instance, butterfly must satisfy Eqs. (2.2), (2.5) and (2.6)
as well as Eq. (2.7).
Umbilic is
As is easily checked this singularity is not realized for the non-transparent
φαβ = 0 .
(2.8)
model.
Beak-to-beak and lips are characterized by
∂zD = 0 ,
a) Lips ∆ ≡ −detDαβ > 0
b) Beak-to-beak ∆ < 0 .
The explicit form of Eq.(2.9) is
∆ takes the form
∂3φ
∂z3 = 0 .
∆ = −4
∂4φ
∂z4 < 0
for non-transparent model. Therefore Eq.(2.12) leads to beak-to-beak.
4
(2.9)
(2.10)
(2.11)
(2.12)
(2.13)
3 Multipole Expansion (MPE) Model
Fermat's potential is
φ =
1
2z − zs2 −
1
2π Z d2x′ κ(x′) lnz − z′2 .
Using the expansion form of logarithmic function,
lnz − z′2 = lnz2 + ln1 −
1
= lnz2 − Xn=1
z′
z 2
nn(cid:16) z′
z (cid:17)n
+ c.c.o ,
we obtain the multipole expansion form of Fermat's potential,
1
2 Xn=1(cid:16) qn
zn + c.c.(cid:17) ,
φ =
m ≡
qn ≡
lnz2 +
m
2
1
2z − zs2 −
1
π Z d2x κ(x) ,
πn Z d2x κ(x) zn .
1
(3.1)
(3.2)
(3.3)
(3.4)
(3.5)
Thus MPE model becomes a non-transparent model from the convergence
requirement of Taylor expansion. This property seems to have interrupted pop-
ularity in gravitational lensing. However this model is very useful and applicable
to many lensing events and deserves further investigations as will be discussed.
We can transform the potential into,
φ ≡
u ≡
φ
m
z
1
2
m
=
,
1
2u − v2 −
zs
m
v ≡
1
2
1
2
,
lnu2 +
1
2 Xn=1(cid:16) Qn
un + c.c.(cid:17) ,
Qn ≡
qn
m1+ n
2
.
For this potential, critical condition and cusp condition become
u2 = 1 + X βn
Qn
un ,
¯Qn
¯un (cid:17)2(cid:16)1 + X βn
¯Qn
¯un (cid:17)2(cid:16)1 + X βn
Imh(cid:16)1 + X αnβn
Reh(cid:16)1 + X αnβn
Qn
un (cid:17)3i = 0 ,
un (cid:17)3i > 0 ,
Qn
(3.6)
(3.7)
(3.8)
(3.9)
(3.10)
where αn ≡ (n + 2)/2 and βn ≡ n(n + 1) . Im (Re) denotes the imaginary
(real) part.
5
4 The Properties of 2n- Pole
Unfortunately, if we consider the terms up to quadrupole moment, MPE model
is already beyond analytical calculation. However, analytical survey is desirable
for the essential physical interpretation. So we consider the case of 2n-pole only,
which allows analytical survey.
Introducing the variables,
= √Xeinθ ,
Y = √X cos nθ ,
un
Qn
Eqs.(3.8), (3.9) and (3.10) are rewritten into,
4
n X 1+ 2
Qn
n = X + 2βnY + β2
n ,
Im(cusp) ≡ sin nθ[(2αn − 3)X 2 + 2βn(α2
n − 3)XY
n(2Y + αnβn)2] = 0 ,
n + 3)XY 2 − (α2
n − 6αn + 1)X − β2
n{2(α2
n(3α2
+β2
Re(cusp) ≡ X 3 + (2αn + 3)βnX 2Y + β2
+3β3
n(α2
n + 2αn − 1)XY + αnβ4
(4.1)
(4.2)
n − 6αn + 3)X 2}
nY (2Y + αnβn)2 > 0 , (4.3)
(4.4)
n(3αn − 2)X + β3
X ≥ Y 2 .
The last condition follows from the just definition of X and Y . As for the
solution of (X, Y ), there are 2n solutions for X > Y 2 on lens plane and n solutions
for X = Y 2.
At first, we examine the cusp conditions, Eqs.(4.2) and (4.3), and Eq.(4.4).
When sin nθ = 0, i.e., X = Y 2, Eq.(4.3) becomes
Re(cusp) = Y (Y + αnβn)2(Y + βn)3 > 0 .
Therefore a part of region that satisfies the cusp conditions is X = Y 2 for Y > 0 or
Y < −βn (Y 6= −αnβn). Further at the points, (X, Y ) = (β2
n,−αnβn),
the beak-to-beak condition is held.
n,−βn) , (α2
nβ2
When sin nθ 6= 0, Eq.(4.2) becomes
(2αn−3)X 2 +2βn(α2
n−3)XY +β2
n(3α2
n−6αn +1)X−β2
n(2Y +αnβn)2 = 0 . (4.5)
After some calculations, we can obtain the region that satisfies Eq.(4.3) and
Eq.(4.4),
X > 0 ,
αnβn
2αn − 3 ≥ Y > −βn .
We call the characteristic points, P1, P2 and P3 defined by
(4.6)
6
P1 = (β2
n, − βn) ,
α2
nβ2
P3 = (cid:16)
n
(2αn − 3)2 ,
nβ2
αnβn
2αn − 3(cid:17) .
P2 = (α2
n, − αnβn) ,
The region in which the cusp conditions is satisfied is depicted in Fig.1 for
n = 2.
-- -- -- --
Fig.1
-- -- -- --
For the other values of n, topological structure does not change.
Next we examine the qualitative behaviour of the critical condition, Eq.(4.1).
Firstly, irrespective of the value of Qn, the curve Eq.(4.1) passes through the
definite point,
(0 , −
βn
2
) ,
and passes above P1,
Y (X = β2
n) = −βn + Qn4/n β1+4/n
2
n
> −βn .
For X > 0, the curve has a locally minimum point and Y tends to ∞ at X = ∞.
Furthermore, for fixed X, the value of Y increases as Qn does.
From the above arguments, the number of the solutions of the critical condi-
tion and the cusp conditions changes when the curve Eq.(4.1)
case (0) is tangential to X = Y 2 for Y < −βn or,
case (1) passes through P3 , or
case (2) is tangential to Im(cusp) = 0 for Eq.(4.6).
Each case is depicted in Fig.2 for n = 2, 4.
-- -- -- -- --
Fig.2(a)
-- -- -- -- --
-- -- -- -- --
Fig.2(b)
-- -- -- -- --
7
For the case(0), the value of Qn is
Qn = Q(0)
n ≡
2
n(n + 1)(n + 2)(cid:16) n
n + 2(cid:17)
n
2 ,
(4.7)
and the tangential point is P2. Therefore beak-to-beak singularity appears.
For the case(1), the value of Qn is
2αn − 3
αnβn (cid:16)3
2 = 3n/2(n − 1)Q(0)
n .
Qn = Q(1)
αn − 1
αn
n ≡
(cid:17)
n
(4.8)
As known from graphical consideration, then cusps split into more cusps or
merge to less cusps. Therefore butterfly singularity appears.
For the case(2), only when n ≤ 3, the lines of Eqs.(4.1) and (4.2) are tangential
to each other in the region Eq.(4.6) and the value of Qn is
Qn = Q(2)
n
≡
qn2 + 9n + 16 + n√n2 + 10n + 17
2n(n + 1)(n + 2)
× nn(n + 1 + √n2 + 10n + 17)
4(n + 2)
on/2
.
(4.9)
And then cusps merge to fold, i.e., swallowtail singularity appears.
We surmmarize the result:
(I) for n = 1, P3 = (∞ , ∞)
(i) Qn < Q(0)
n
There are three solutions on X = Y 2 line and two solutions for X >
Y 2, so 3 + 2 × 2 = 7 cusps on lens plane.
(ii) Q(0)
n < Qn < Q(2)
n
After beak-to-beak singularity appears, there is a solution on X = Y 2
line and two solutions for X > Y 2, so 1 + 2 × 2 = 5 cusps on lens
plane.
(iii) Q(2)
After swallowtail singularity appears, there is a solution on X = Y 2
line and no solution for X > Y 2, so one cusp on lens plane.
n < Qn
(II) for n = 2, 3
(i) Qn < Q(0)
n
There are three solutions on X = Y 2 line and one solution for X > Y 2,
so 3n + 2n = 5n cusps on lens plane.
(ii) Q(0)
n < Qn < Q(1)
n
After beak-to-beak singularity appears, there is a solution on X = Y 2
line and one solution for X > Y 2, so n + 2n = 3n cusps on lens plane.
8
(iii) Q(1)
n < Qn < Q(2)
n
After butterfly singularity appears, there is a solution on X = Y 2 line
and two solutions for X > Y 2, so n + 2× 2n = 5n cusps on lens plane.
(iv) Q(2)
n < Qn
After swallowtail singularity appears, there is a solution on X = Y 2
line and no solution for X > Y 2, so n cusps on lens plane.
(III) for n ≥ 4
(i) Qn < Q(0)
n
There are three solutions on X = Y 2 line and one solution for X > Y 2,
so 3n + 2n = 5n cusps on lens plane.
(ii) Q(0)
n < Qn < Q(1)
n
After beak-to-beak singularity appears, there is a solution on X = Y 2
line and one solution for X > Y 2, so n + 2n = 3n cusps on lens plane.
(iii) Q(1)
n < Qn
After butterfly singularity appears, there is a solution on X = Y 2 line
and no solution for X > Y 2, so n cusps on lens plane.
5 Two Point-Mass Model
Critical curve produced by the lens galaxy of PG1115+080 makes us imagine
that the lens galaxy may be approximated by the two point-mass lens [Fig.3].
-- -- -- --
Fig.3
-- -- -- --
This model is interesting as its own right from the following reason. In the
previous section we studied the complex lens object by decomposing it into 2n-
pole contribution. Two point-mass model is very simple one but is composed of
composite multipoles.
So in this section we analyse this model in the analogous manner as in the
previous section. It is found that this model is qualitatively different from the
MPE model.
This model has been discussed by Erdl and Schneider [8]. A straightforward
application of catastrophe theory to this model still enforces numerical calcula-
tions. However we can derive singularities in simple analytical ways by pursuing
the critical conditions of f (r, ρ) defined in Eq.(5.8).
The surface mass density of two point-mass lens is given by
Σ(~x) = M[µAδ2(~x − ~a) + µBδ2(~x + ~a)]
,
(5.1)
9
masses in the lens surface.
here M is total mass and µA + µB = 1. ~a and −~a are locations of two point
Fermat potential φ is given by
φ =
1
2z − zs2 −
m
2
(µA lnz − χ2 + µB lnz + χ2) ,
in complex representation with m ≡ M
We can rescale the potential into
πΣcr
and χ ≡ a1 + ia2.
φ
m
1
2u − v2 −
1
2
=
φ ≡
[µA lnu − ξ2 + µB lnu + ξ2]
up to irrelevant constant. Here u ≡ z√m, v ≡ zs√m and ξ ≡ χ√m.
Lens equation ∂φ
∂u = 0 gives
(5.2)
(5.3)
Critical curve which is given by D = 1
coordinates takes the following form,
¯v = ¯u − (
µA
u − ξ
+
) .
µB
u + ξ
4 − φuu2 = 0 in general system of
(5.4)
µA
(u − ξ)2 +
µB
(u + ξ)2 = 1 .
(5.5)
Introducing the new complex variable U ≡ u+ξ
ξ , ξ in Eq.(5.5) is factorized out
as
ξ2 = U 2 − 4µBU + 4µB
U(U − 2)2
.
(5.6)
Using the polar coordinates U = reiθ, critical curve Eq.(5.6) is rewritten as
where
ξ4 = f (r, cos θ) ,
16µBr2ρ2 − 8µBr(r2 + 4µB)ρ + r4 + 8µB(2µB − 1)r2 + 16µ2
B
f (r, ρ) ≡
with ρ ≡ cos θ.
Beak-to-beak condition ∂D
r4(r2 − 4rρ + 4)2
∂U = 0 becomes
from Eq.(5.7). Eq.(5.9) is equivalent to
∂f
∂U
= 0
Re(
∂f
∂U
) = ρ
∂f
∂r
+
1 − ρ2
r
∂f
∂ρ
= 0
10
(5.7)
(5.8)
(5.9)
(5.10)
and
and, therefore, to
for sin θ 6= 0 or to
Im(
∂f
∂U
) = sin θ(
∂f
∂r −
ρ
r
∂f
∂ρ
) = 0 .
case i)
∂f
∂r
=
∂f
∂ρ
= 0
case ii)
sin θ =
∂f
∂r
= 0 .
(5.11)
(5.12)
(5.13)
Beak-to-beak appears at the points where four real parameters r, ρ, µB and
ξ satisfy three equations (5.7) and (5.12) or (5.13).
of f (r, ρ).
Thus metamorphoses of the caustics are appeared in the critical behaviours
Firstly we consider case i).
∂f
∂ρ
=
8(1 − µB)(16µBρ − 12µBr + r3)
r2(r2 − 4rρ + 4)3
.
(5.14)
From the definitions, 1 − µB > 0, r2 − 4rρ + 4 > 0 (ρ < 1 for case i)) and f
takes the minimum value Fm(r) at
ρ =
3
4
r −
r3
16µB ≡ m(r)
(5.15)
for fixed r.
Fm(r) ≡ f (r, m(r)) =
r4(r4 − 8µBr2 + 16µB)
We proceed to study the critical behaviour of Fm(r).
µB(r2 − 4µB)2
(> 0) .
(5.16)
∂Fm(r)
∂r
= −
4µB(r2 − 4µB)
r5(r4 − 8µBr2 + 16µ2
B)
g(r2) ,
where
(5.17)
(5.18)
g(r2) ≡ r6 − 12µBr4 + 48µ2
Br2 − 64µ2
B .
g(r2) is a monotonically increasing function of r2 and the root of g(r2) = 0 is
given by r2
1,
1 = 4µ2/3
r2
B (µ1/3
A + µ1/3
B ) ≡ 4µ2/3
B δ .
(5.19)
The behaviour of Fm(r) is depicted in Fig.4.
11
-- -- -- -- -
Fig.4
-- -- -- -- -
Nextly we consider case ii) where ρ = ±1. We define f± ≡ f (r,±1).
Then
f− =
(r2 + 4µBr + 4µB)2
r4(r + 2)4
is a monotonically decreasing function of r and has no critical point. On the
other hand,
and
where
f+ =
(r2 − 4µBr + 4µB)2
r4(r − 2)4
df+
dr
= −
4
r5(r − 2)5 (r2 − 4µBr + 4µB) h(r) ,
h(r) ≡ r3 − 6µBr2 + 12µBr − 8µB .
(5.20)
(5.21)
(5.22)
h(r) is a monotonically increasing function of r and has only one real solution
r2 to h(r) = 0,
r2 = 2µ1/3
B (µ2/3
A + µ2/3
B − µ1/3
A µ1/3
B ) =
2µ1/3
B
δ
.
(5.23)
The behaviour of f+ is depicted in Fig.5.
-- -- -- -- -
Fig.5
-- -- -- -- -
The above discussions in case i) and case ii) together with the facts that
f (0, ρ) = ∞ , f (∞, ρ) = 0 and
f (2, 1) = ∞
(5.24)
are summarized as the contour map of f (r, ρ) [Fig.6].
-- -- -- -- -
Fig.6
-- -- -- -- -
12
It is found that the critical curve changes its topology when we cross the
contour L1 (L2) passing through C1 (C2). We denote the ξ values corresponding
to the contour L1 (L2) as ξ1 (ξ2). Then
and
ξ14 = fC1 =
1
16δ3
ξ24 = fC2 =
δ6
16
,
(5.25)
(5.26)
which agree with the results of Erdl and Schneider [8].
6 The Application to the Multiple Quasar PG1115+080
In this section, we apply our MPE model to the lensing galaxy of multiple quasar
PG1115+080. We use deflection potential up to the 23-pole expansion here.
About this lensed quasar, we have got the observed data of positions and
relative amplifications of the images [9]. (We have adopted the new data for only
zl from Angonim-Willaime et al. [10]) The positions of the images are as follows;
θ(A1) = (−0′′.94,−0′′.73)
θ(A2) = (−1′′.11,−0′′.27)
θ(B) = (0′′.72,−0′′.60)
θ(C) = (0′′.33, 1′′.35)
And the relative amplifications of the images are as follows;
µ(A1)/µ(C) = 3.22
µ(A2)/µ(C) = 2.49
µ(B)/µ(C) = 0.64
Lens equation up to the 23-pole expansion is given by
v∗ − u∗ +
1
u
+
D
u2 +
2Q
u3 +
3T
u4 = 0 .
(6.1)
(6.2)
(6.3)
Here v is source position, u is image position, and D, Q, T are dipole,
quadrupole and 23-pole, respectively. These are all dimensionless complex num-
bered quantities normalized by suitable power of mass (See Eq.(3.7)).
Image
amplification is given by
13
µ = 1 −
m
z2 +
2d
z3 +
6q
z4 +
12t
z5 2−1 .
(6.4)
We obtain values of the parameters included in Eq.(6.3) which reproduce the
observed data well. These parameter values are as follows;
source position = (0′′.02785, 0′′.01507)
[rad2]
m = 3.496 × 10−11
D = 0.1824 ,
Q = 0.08724 ,
T = 3.407 × 10−3 ,
θD = 88◦.81
θQ = 130◦.6
θT = 115◦.5
(6.5)
If we assume Einstein-de Sitter universe, this mass would be M = πΣcrD2
l m =
1.483± 0.024 h−1 × 1011 M⊙. Here redshifts of the lensing galaxy and the source
quasar are zl = 0.294 ± 0.005 and zs = 1.722, respectively. Hubble constant is
H0 = 100 h km/sec/Mpc.
Calculated image positions are
θ(A1) = (−0′′.9383,−0′′.7317)
θ(A2) = (−1′′.110,−0′′.2748)
θ(B) = (0′′.7205,−0′′.6004)
θ(C) = (0′′.3294, 1′′.350)
θ(D) = (0′′.5915,−0′′.4450)
θ(E) = (−0′′.4494, 0′′.08641) .
And calculated relative amplifications are
µ(A1)/µ(C) = 3.113
µ(A2)/µ(C) = 2.892
µ(B)/µ(C) = 0.7467
µ(D)/µ(C) = 0.1780
µ(E)/µ(C) = 0.005871 .
(6.6)
(6.7)
Then, we should check the validity of approximation in multipole expansion
up to 23-pole. It must hold the following conditions.
1 ≫ D
u ≫ Q
u2 ≫ T
u3
And the best fit parameters (6.6) give following values.
(6.8)
14
T
u3 ≈ 1.621 × 10−2
u3 . So this MPE model up to
(6.9)
u2 ≫ T
Q
u2 ≈ 0.2103 ,
This result roughly holds the condition Q
D
u ≈ 0.2609 ,
23-pole has been proved to be valid up to 23-pole moment.
We performed parameter fitting of this lensed quasar using MPE model up to
quadrupole in the previous article [7]. The values of source position, dipole and
quadrupole in Eq.(6.6) are similar to those values in previous fitting. So they are
consistent.
7 The Relations with Other Models
We have shown in the previous sections that deflection potential up to the
quadrupole expansion reproduces multiple images well. Its critical curve takes
the similar form produced by two point-mass lens [8].
So it may be meaningful to check that two point-mass lens model on the single
lens plane well resembles with the lens of PG1115+080.
Also it is very helpful for the study of lensing object that we analyze the
same lensing system by qualitatively different models. As a typical example, we
consider softened power-law spheroid models.
7.1 Two point-mass lens
Fermat potential given by Eq. (5.2) is expanded by multipoles as
φ =
1
2~x − ~y2 − m ln~x −
dixi
~x2 −
qijxixj
~x4 − ...
.
(7.1)
Here
M
m ≡
πΣcr
di ≡ mXk
qij ≡ mXk
µkdki
(k = A, B)
(7.2)
µk(dkidkj − ~dk2
2
δij)
and so on, which are the real representations of Eqs.(3.3) - (3.5).
In this case we have six independent parameters M , µA , ~dA , ~dB besides source
positions ~y. Here we assume that MPE model up to quadrupole moments is good
approximation as we have shown and that two point-mass lens can be considered
to be their prototype. Then we must be able to fit the parameters M , µA , ~dA , ~dB
15
from Eq.(7.2) so as to give the same m , di , qij as the MPE model. This is always
possible. For Eq.(7.2) consists of five equations and we have six parameters to
be adjusted.
Whether two point-mass model is good approximation or not, therefore, de-
pends on the fact that the higher multipole terms subsequent to quadrupole in
Eq.(7.1) is small or not. Here higher multipole terms is fixed by M , µA , ~dA , ~dB
determined by Eq.(7.2).
As is easily checked it is impossible. Therefore two point-mass lens model
cannot be a good approximation of PG1115+080.
7.2 Softened power-law spheroids (SPLS)
This model is also called elliptical lens model whose mass density is given by
ρ(a) = ρ0(1 +
)− k
2
a2
r2
c
(7.3)
(7.4)
with
x′2 + y′2 +
z′2
1 − e2 = a2
The principal axes of spheroid (x′, y′, z′) are related with the coordinate of
optical axis and lens surface (x, y, z) by
x′
y′
z′
=
0
0
1
0 cos γ − sin γ
cos γ
0 sin γ
x
y
z
.
(7.5)
That is, optical axis z is tilted to z′ by γ.
Here rc is the core radius and ρ0 is the constant central density. e is the
eccentricity. For k ≤ 3 total mass is divergent, mass density has a cut-off
ρ(a) = 0
for
a
rc
> n .
(7.6)
This model has been applied to PG1115+080 by Narasimha et al. [11]. How-
ever they used the observational data different from ours, therefore we cannot
simply compare their results with ours. Elliptical lens model is qualitatively
different model from MPE model. The former is transparent and the latter non-
transparent. If we are only satisfied by reproducing image positions, both model
may pass our requirements. So we must try parameter fitting to more various
and precise data.
The detail of this model has been discussed in a separate form [6], so we just
quote the result [Table 1, Table 2].
16
-- -- -- -- --
Table 1
-- -- -- -- --
-- -- -- -- --
Table 2
-- -- -- -- --
SPLS has a large fitting zone for its transparency compared with MPE.
As a whole this model gives rather non-compact lensing object unlike the
compact aspect of Narasimha et al. [11]. This seems to be incompatible with the
observed lensing object [12]. Observed image amplifications are better fitted by
MPE model than by elliptical lens model.
8 Discussions
MPE model is not so popular as the elliptical lens model with or without external
shear since most of the gravitational lensings are thought to be caused by the
transparent global systems composed of many galaxies. However, the recent
observations support that the lens of PG1115+080 is a compact object whose
optical extent is within the observed innermost image B ∼ 10h−1 kpc [9] [12].
MPE model is neither transparent nor nonsingular. Therefore it is free from the
general theorems about gravitational lensing (See chapter 5.4 of Schneider et al.
[1]). Indeed, in Fig.3 we have 4 images of odd parity and 2 images of even parity.
There gives rise, of course, no inconsistency.
We must wait further refinement of model check in order that our model
parameters are physically affirmative. For that purpose, we need more precise
observations such as image shape, time delays and spectroscopy profile [13], etc.
In this connection it is in order to comment on time delay.
Time delay is given by
t =
1
c
(1 + zl)
DlDs
Dls
φ .
(8.1)
For Ω0 = 1 case,
Dij =
(
For Ω0 = 0 case,
2c
H0
zi < zj ,
(1 + zj)√1 + zi − (1 + zi)√1 + zj
(1 + zj)2(1 + zi)
i, j = source, lens, observer )
.
(8.2)
Dij =
c
2H0
(1 + zj)2 − (1 + zi)2
(1 + zj)2(1 + zi)
.
(8.3)
17
Time delays for each case are given in Table 3.
-- -- -- -- --
Table 3
-- -- -- -- --
For comparison we have also listed the respective time delay based on the
elliptical lens model [Table 2]. Estimations by MPE model are roughly coinci-
dent with the observation by Vanderriest et al.[14]. On the contrary, estimations
by SPLS in Table 2 are too short, which denies the possibility that its amplifi-
cation discrepancy with observation might be remedied by the time variation of
amplification of source QSO. Naively we may conclude that the lensing object of
PG1115+080 is compact also from the theoretical side.
Consequently MPE model deserves serious surveys.
In this connection it
should be remarked that the dipole term plays an important role in our theory :
The presence of dipole moment means that the origin of the coordinates (which
is the optical center) is away from the center of mass of the lensing galaxy.
There are the arguments which insist that this deviation may be the cause of
spiral arms. Also MPE model gives better fit to the observational data than
elliptical lens model in many points [6]. This may supports the view point that
the lensing galaxy is a spiral galaxy, whose equipotential contour is different from
spheroid. There is the observational support of this viewpoint from the analysis
of spectroscopy [15][13]. From these facts, non zero dipole component may play
any role in the formation of spiral arms.
Though we applied it only to PG1115+080 in this article, MPE model is
very useful for the wide class of lensing phenomena. It gives systematically the
multipole moments of the lensing object, irrespective to isolated lens or not. This
serves crucially to the construction of more concrete and more precise structure
of lensing object.
Acknowledgements
We are grateful to Dr. Hiroshi Yoshida for useful comments.
18
References
[1] P. Schneider, J. Ehlers and E. E. Falco, Gravitational Lenses (Springer-
Verlag, 1992).
[2] R. Blandford and C. S. Kochanek, "Gravitatinal Lenses" in Dark Matters in
the Universe, ed. by J. Bahcall, T. Piran and S. Weinberg (World Scientific,
1987).
[3] E. Roulet and S. Mollerach, "Microlensing" (astro-ph/9603119) (1996) -- to
appear in Physics Reports.
[4] J. Binney and S. Tremain, "Galactic Dynamics" (Princeton University Press,
1987).
[5] "Digital Sky Survey of the Northern Galactic Cap, ed. by J. E. Gunn (Prince-
ton Univ., 1994).
[6] K. Asano and T. Fukuyama, Preprint Ritsumei-pp-12 (Revised) "Elliptical
Lens Model in PG1115+080" (1996).
[7] Y. Kakigi, T. Okamura and T. Fukuyama,
Intern. J. Mod. Phys.
D 4, 685 (1995).
[8] H. Erdl and P. Schneider, Astron. Astrophys. 268, 453 (1993).
[9] C. A. Christian, D. Crabtree and P. Waddell, Ap. J. 312, 45 (1987).
[10] M. C. Angonim-Willaime, F. Hammer and F. Rigaut, in Proc. 31'st Liege Int.
Astrophys. Collog., Gravitational Lenses in the Universe, ed. by J. Surdej et
al. (Liege Univ. Liege), 85 (1993).
[11] D. Narasimha, K. Subramanian and S. M. Chitre, Mon. Not. R. astr.
Soc. 200, 941 (1982).
[12] J. Kristian et al., A. J. 106, 1330 (1993).
[13] A. G. Michalitsianos, R. J. Oliversen and J. Nichols, Ap. J. 461, 593 (1996).
[14] C. Vanderriest, G. Wl´erick, G. Leli`evre, J. Schneider, H. Sol, D. Harville, L.
Renard and B. Servan, Astron. Astrophys. 159, L5 (1986).
[15] E. L. Turner, Proc. 14th Texas Symp. on Relativistic Astrophysics, ed. by E.
Fenyves (New York, NY Aead. Sci), 319 (1989).
19
Figure Captions
Fig.1 The region that satisfies the cusp condition. Solid line is X = Y 2, dotted
line is Re(cusp)=0 and dashed line is Im(cusp)=0. Three lines pass through
the points P1 and P2. The cusp condition is satisfied in the region where
the dashed line is in the common region of the interiors of the solid and the
dotted line. Between P1 and P2, the dashed line passes very close to but in
the exterior of the dotted line.
Fig.2 The behaviours of the critical condition. Solid line is X = Y 2, dashed
line is Im(cusp)=0 and dotted line is the critical condition. Fig.2(a) is the
n = 2 case. The dotted lines of (0), (1) and (2) correspond to the cases (0),
(1) and (2) discussed in the text, respectively. Fig.2(b) is the n = 4 case.
Fig.3 Shapes of images, critical line and caustics for the multiple quasar PG1115+080
(Kakigi et al. 1995). Crosses (dotted regions) are the observed (calculated)
image positions. It has been assumed that source QSO is spherical.
Fig.4 The critical behaviour of Fm(r). Critical points C0 and C1 in (r, ρ) plane
are (r0, ρ0) ≡ (2µ1/2
A )),
respectively. We can set µB ≥ µA without spoiling generality and therefore
1 > ρ1 > 0.
B ) and (r1, ρ1) ≡ (2µ1/3
B − µ1/3
B , µ1/2
B δ1/2, δ1/2
2 (2µ1/3
Fig.5 The critical behaviour of f+. Critical point C2 is given by C2 = (r2, ρ2) =
( 2µ1/3
B
δ
, 1).
Fig.6 The contour map of f (r, ρ). Origin is the point of heavier point mass.
f (r, ρ) diminishes in the direction of ⇒. C1 and C2 are the points where
beak-to-beaks appear.
Table 1 Fitting parameters of SPLS ; (a) k=3 case and (b) k=2 case.
Table 2 Calculated image positions (and time delays lagging behind C in h−1
days) of SPLS ; (a) k=3 case and (b) k=2 case.
Table 3 Time delays lagging behind C in h−1 days of MPE model.
20
0
r
dFm
dr
C1
r1
0
−
Fm ∞ ց 0 ր 1
16δ3 ց
C0
2µ1/2
B
0
−
+
Fig.4
0
r
df+
dr
C2
r2
− 0
f+ ∞ ց δ6
Fig.5
2
−
16 ր ∞ ց
e sin γ
0.4
0.4
0.5
0.5
0.6
0.6
e sin γ
0.3
0.3
0.4
0.4
0.5
0.5
n
5
10
5
10
5
10
n
5
10
5
10
5
10
rc
6.13 h−1 kpc
6.07 h−1 kpc
4.78 h−1 kpc
4.71 h−1 kpc
3.84 h−1 kpc
3.75 h−1 kpc
rc
6.00 h−1 kpc
5.46 h−1 kpc
4.23 h−1 kpc
3.97 h−1 kpc
3.13 h−1 kpc
2.84 h−1 kpc
κ0√1 − e2
2.58×10−23 h2g/cm3
2.56×10−23 h2g/cm3
3.48×10−23 h2g/cm3
3.50×10−23 h2g/cm3
4.66×10−23 h2g/cm3
4.74×10−23 h2g/cm3
Table 1 (a)
κ0√1 − e2
1.79×10−23 h2g/cm3
1.87×10−23 h2g/cm3
2.71×10−23 h2g/cm3
2.71×10−23 h2g/cm3
3.93×10−23 h2g/cm3
4.13×10−23 h2g/cm3
Table 1 (b)
Total Mass
Source Position
1.48×1012M⊙ (0′′.024, − 0′′.037)
2.13×1012M⊙ (0′′.024, − 0′′.037)
9.43×1011M⊙ (0′′.038, − 0′′.058)
1.37×1012M⊙ (0′′.039, − 0′′.058)
6.56×1011M⊙ (0′′.057, − 0′′.085)
9.30×1011M⊙ (0′′.057, − 0′′.084)
Total Mass
Source Position
2.62×1012M⊙ (0′′.014, − 0′′.020)
4.83×1012M⊙ (0′′.014, − 0′′.021)
1.39×1012M⊙ (0′′.025, − 0′′.036)
2.70×1012M⊙ (0′′.025, − 0′′.035)
8.15×1011M⊙ (0′′.039, − 0′′.055)
1.51×1012M⊙ (0′′.040, − 0′′.054)
e sin γ
0.4
0.4
0.5
0.5
0.6
0.6
e sin γ
0.3
0.3
0.4
0.4
0.5
0.5
n
5
10
5
10
5
10
n
5
10
5
10
5
10
A1
A2
B
C
(0′′.51, 1′′.08) (4.0)
(0′′.51, 1′′.08) (3.9)
(0′′.53, 1′′.06) (6.1)
(0′′.53, 1′′.06) (6.2)
(0′′.53, 1′′.05) (9.0)
(0′′.53, 1′′.05) (8.8)
(0′′.95, 0′′.63) (4.1)
(0′′.96, 0′′.62) (4.0)
(0′′.97, 0′′.61) (6.3)
(0′′.97, 0′′.61) (6.3)
(1′′.00, 0′′.57) (9.2)
(1′′.00, 0′′.57) (9.1)
(−0′′.88, 0′′.31) (6.0)
(−0′′.89, 0′′.31) (5.9)
(−0′′.89, 0′′.30) (9.4)
(−0′′.89, 0′′.30) (9.6)
(−0′′.89, 0′′.28) (13.9)
(−0′′.89, 0′′.28) (13.8)
(0′′.23, − 1′′.37)
(0′′.24, − 1′′.37)
(0′′.24, − 1′′.37)
(0′′.25, − 1′′.37)
(0′′.26, − 1′′.37)
(0′′.26, − 1′′.37)
Table 2 (a)
A1
A2
B
C
(0′′.52, 1′′.08) (2.1)
(0′′.56, 1′′.04) (2.2)
(0′′.54, 1′′.06) (3.9)
(0′′.52, 1′′.07) (3.8)
(0′′.54, 1′′.05) (5.8)
(0′′.54, 1′′.05) (5.8)
(0′′.94, 0′′.64) (2.1)
(0′′.92, 0′′.67) (2.3)
(0′′.96, 0′′.62) (4.0)
(0′′.92, 0′′.61) (4.0)
(0′′.99, 0′′.58) (6.0)
(0′′.99, 0′′.57) (6.0)
(−0′′.88, 0′′.32) (3.2)
(−0′′.90, 0′′.31) (3.4)
(−0′′.88, 0′′.31) (6.1)
(−0′′.89, 0′′.30) (6.1)
(−0′′.89, 0′′.29) (9.3)
(−0′′.89, 0′′.29) (9.4)
(0′′.23, − 1′′.37)
(0′′.23, − 1′′.35)
(0′′.26, − 1′′.36)
(0′′.26, − 1′′.36)
(0′′.28, − 1′′.36)
(0′′.28, − 1′′.36)
Table 2 (b)
∆ (h−1 days)
Ω0 = 1
Ω0 = 0
C − A1
9.757 ± 0.200
10.540+0.227
−0.225
C − A2
11.448 ± 0.234
12.367+0.266
−0.264
C − B
19.407 ± 0.396
20.964+0.453
−0.447
C − D
19.225 ± 0.392
20.767+0.448
−0.443
C − E
3.076 ± 0.063
3.323+0.072
−0.071
Table 3
|
astro-ph/0608257 | 1 | 0608 | 2006-08-11T20:01:22 | The gamma-ray burst GRB060614 requires a novel explosive process | [
"astro-ph"
] | Over the past decade our physical understanding of gamma-ray bursts (GRBs) has progressed rapidly thanks to the discovery and observation of their long-lived afterglow emission. Long-duration (T < 2 s) GRBs are associated with the explosive deaths of massive stars (``collapsars''), which produce accompanying supernovae, while the short-duration (T > 2 s) GRBs arise from a different origin, which has been argued to be the merger of two compact objects, either neutron stars or black holes. Here we present observations of GRB060614, a 100-s long burst discovered by the Swift satellite, which require the invocation of a new explosive process: either a massive ``collapsar'' that powers a GRB without any associated supernova, or a new type of engine, as long-lived as the collapsar but without any such massive stellar host. We also discuss the properties of this burst's redshift z=0.125 host galaxy, which distinguish it from other long-duration GRBs and suggest that an entirely new type of GRB progenitor may be required. | astro-ph | astro-ph |
The γ-ray burst GRB060614 requires a novel explosive
process
A. Gal-Yam,1 D. B. Fox,2 P. A. Price,3 M. R. Davis,4 D. C. Leonard,4 A. M. Soderberg,1
E. Nakar,1 E. O. Ofek,1 B. P. Schmidt,5 K. Lewis,5 B. A. Peterson,5 S. R. Kulkarni,1
E. Berger,6 S. B. Cenko,1 R. Sari,1 K. Sharon,7 D. Frail,8 N. Gehrels,9 J. A. Nousek,2
D. N. Burrows,2 V. Mangano,10 S. T. Holland,9 P. J. Brown,2 D.-S. Moon,1 F. Harrison,1
T. Piran10 S. E. Persson,6 P. J. McCarthy,6 B. E. Penprase,11 & R. A. Chevalier12
1 Division of Physics, Mathematics and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA
2 Department of Astronomy and Astrophysics, Pennsylvania State University, 525 Davey Lab, University Park, PA 16802,
USA
3 University of Hawaii, Institute of Astronomy, 2680 Woodlawn Drive, Honolulu, HI 96822-1897, USA
4 Department of Astronomy, San Diego State University, San Diego, California 92182, USA
5 Research School of Astronomy and Astrophysics, Australian National University, Mt Stromlo Observatory, via Cotter
Rd, Weston Creek, ACT 2611, Australia
6
Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101, and Princeton
University Observatory, Peyton Hall, Ivy Lane, Princeton, NJ 08544, USA
7 Department of Astrophysics, Tel Aviv University, 69978 Tel Aviv, Israel
8 National Radio Astronomy Observatory, P.O. Box 0, Socorro, New Mexico 87801, USA
9 NASA/Goddard Space Flight Center, Greenbelt, Maryland 20771, USA
10 INAF, Istituto di Astrofisica Spaziale e Fisica Cosmica di Palermo, Via Ugo La Malfa 153, I-90146, Palermo, Italy
10 Racah Institute of Physics, Hebrew University, Jerusalem 91904, Israel
11 Pomona College Dept. of Physics & Astronomy, 610 N. College Ave, Claremont, CA 91711, USA
12 Department of Astronomy, University of Virginia, PO Box 3818, Charlottesville, VA 22903, USA
Over the past decade our physical understanding of gamma-ray bursts (GRBs) has progressed
rapidly thanks to the discovery and observation of their long-lived afterglow emission. Long-
duration (T >
∼ 2 s) GRBs are associated with the explosive deaths of massive stars ("collap-
sars"1), which produce accompanying supernovae,2 -- 4 while the short-duration (T <
∼ 2 s) GRBs
arise from a different origin, which has been argued to be the merger of two compact objects,5 -- 7
either neutron stars or black holes. Here we present observations of GRB 060614, a 100-s long
burst discovered by the Swift satellite,8 which require the invocation of a new explosive process:
2
Gal-Yam et al.
either a massive "collapsar" that powers a GRB without any associated supernova, or a new
type of engine, as long-lived as the collapsar but without any such massive stellar host. We also
discuss the properties of this burst's redshift z = 0.125 host galaxy, which distinguish it from
other long-duration GRBs and suggest that an entirely new type of GRB progenitor may be
required.
On 14 June 2006, 12:43 UT, the burst alert telescope (BAT) on board the Swift satellite
detected the γ-Ray Burst (GRB) 060614.8 The BAT detected γ-rays from this event for
120s, and the burst showed strong variability during much of that period, as confirmed
by parallel observations by the Konus-Wind satellite.9 Note that while some evolution in
the temporal and spectral properties of this GRB were observed, the emission remained
highly variable and relatively hard for tens of seconds, unlike the situation observed for a
few short bursts with long, soft "tails".10,7 This indicates sustained activity of an engine,
rather than the early onset of the afterglow. The γ-ray properties of this event are similar
to those of other bursts from the long-duration subgroup of GRBs. Swift autonomously
slewed to the GRB position and began taking data with the X-ray telescope and UV-
optical telescope.11 We began observing this event ≈ 26 minutes later using the 40 inch
telescope at Siding Springs Observatory. The evolution of the optical radiation from this
event as traced by our data, augmented by Swift observations and additional data from
the literature is shown in Fig. 1. As the optical source decayed, we noticed that it was
apparently superposed on a faint dwarf host galaxy. On June 19, 2006 UT We obtained
a spectrum of the host using the GMOS-S spectrograph mounted on the Gemini-south
8m telescope at Cerro Pachon, Chile. From this spectrum we derived the redshift of the
host galaxy, and by association of the GRB, and found it to be z = 0.125, a low value for
long GRBs. We confirmed this redshift with a higher quality spectrum obtained using the
same instrument on July 15, 2006 UT (Fig. 2). Previous long GRBs at such low redshifts
showed clear signatures of the underlying supernova (SN) explosions at comparable age
post-burst.3,12 However, such signatures were lacking in the case of this long GRB.13
Thus motivated, we undertook target-of-opportunity observations with the Hubble
Space Telescope (HST), combining resources from our approved programs (GO 10551, PI
Kulkarni, GO 10624, 10917, PI Fox). We observed the location of GRB 060614 using the
Wide Field and Planetary Camera 2 (WFPC2) on board HST on June 27-28, 2006 UT, and
again using the Advanced Camera for Surveys (ACS) on July 15-16 2006 UT. Inspection
of the data (Fig. 3) reveals a point source at the outskirts of the GRB host which is
A new explosive origin for GRBs
3
well-detected in our first-epoch WFPC2 observations, and is apparently gone during our
next visit. We identify this object as the optical afterglow of GRB 060614, and derive its
brightness using image-subtraction methods. A decomposition of the measured flux into
the contribution from the GRB afterglow (residual decaying radiation from the interaction
of the GRB ejecta with itself and/or the surrounding material) and that of a possible
supernova (whose optical radiation is dominated by energy released from radioactive decay
of newly synthesized elements, mostly Ni 56), requires detailed modelling of the multiband
evolution of the GRB afterglow which is beyond the scope of this paper. However, our
analysis (Fig. 1) shows that our HST detection is probably dominated by the afterglow,
rather than a possible SN, which is not required by the data. Any putative SN component
must be more than 100 times fainter than the faintest event previously known to be
associated with a long GRB (SN 2006aj/ GRB 06021812,14; Fig. 1).
In fact, such a
supernova would be fainter than any SN ever observed.15 A conservative upper limit on
the amount of synthesized Ni 56 is 8 × 10−4 M⊙, more than two orders of magnitude less
than the typical amount synthesized by long GRB/SNe. Our HST data thus indicate that
this GRB was not associated with a radioactively-powered event similar to any known
SN.
Furthermore, our HST and ground-based data reveal that the properties of the host
of GRB 060614 are unusual when compared to those of the large sample of previously
observed long GRBs. The star formation rate we measure from the spectrum of the host,
0.0035 M⊙ y−1, is very small, and even the specific star formation rate, correcting for
the low luminosity of this dwarf galaxy (L ≈ 0.015L∗) is ∼ 0.23 M⊙ y−1 (L/L∗)−1, a
value which is more than 20 times below that of typical long GRBs16 and more than two
orders of magnitude below that of long GRBs detected in low-luminosity dwarfs at low
redshift.17,18
It is worthwhile to note in this context that unlike long GRBs, members of the short
GRB group have been shown to reside in host galaxies of all types, including elliptical
galaxies with virtually no young, massive stars,6,7,5 and in the outskirts of dwarf galaxies,10
indicating a mature (rather than short-lived) population of progenitors (and no associated
SNe).
Considering the entire set of observations available for this event, the emerging picture
is a puzzling one. On the one hand, the high-energy (γ-ray) properties of this burst are
only consistent with those of a long GRB. On the other hand, the lack of an associated
4
Gal-Yam et al.
SN is inconsistent with an origin in a massive, rapidly-rotating star undergoing a core-
collapse SN explosion ("collapsar"1), the popular, observationally supported model for
long GRBs. Furthermore, the properties of the host galaxy of this event stand out from
among those of numerous other long GRBs observed so far, and rather resemble those of a
short GRB. We now consider possible explanations for this puzzle, and their implications.
First, we can conjecture that all long GRBs result from massive "collapsars", but
that most of these are associated with SNe with a range of properties (as observed so
far19,20) and a minority (the first example of which is GRB 060614) synthesize very little
Ni 56, and do not produce an optically luminous SN. Indeed, there are models of massive
stars that collapse into a black hole and produce, if at all, very faint SNe.21 However,
such events are understood essentially as non-rotating stars, collapsing directly into a
black hole, while "collapsar"-like GRB models generically require a rapidly rotating core,
leading to an accretion disk onto a spinning BH that powers and/or collimates the outgoing
beamed ultra-relativistic ejected material (but see22). The large angular momentum in
such systems requires an accretion disk to facilitate infall into the central BH, and is
thus inconsistent with the non-rotating, direct-infall models for non-SN massive stellar
collapses. In particular, GRB 060614 emitted a large flux (1050 erg s−1) of high-energy
photons and displayed rapid variability, indicating an ultra-relativistic flow with Lorentz
factor of Γ > 15 (calculated following the methodology presented by Ref.
[23]). Such a
flow is not expected in a nearly-spherical direct collapse of a star into a BH, and we thus
consider this option to be apparently at odds with theoretical understanding of massive
stellar death.
Second, especially given the remarkable properties of the host galaxy of GRB 060614,
one might suggest that this is an extreme member of the short GRB group, which are not
associated with massive stellar progenitors, and may result from compact binary mergers.
However, to maintain the putative association with binary mergers would involve a major
revision in our understanding of accretion physics.
In particular, low viscosity during
compact binary mergers are required to allow such a process to continue beyond a fraction
of a second, in contrast with the current consensus.23 Alternative interpretation would
be that this is indeed an extreme member of the short GRB group, and that its existence
indicates that this entire group does not result from compact binary mergers (which can
only produce short duration bursts) but rather from a different physical process altogether,
capable of producing both short and long events, and not associated with young massive
stars. An obvious and testable prediction of these options is that such a "long" member of
the "short" (binary merger or other mechanism) group should be detected in an elliptical
galaxy containing only old stars. Such galaxies host a significant fraction of the "short"
GRB population, but have never so far been associated with events longer than a few
A new explosive origin for GRBs
5
seconds.
Finally, GRB 060614 may be the first example of a new class of GRBs, different from
both typical long events (which are associated with SNe and powered by infall onto a
newly formed BH) and short events (which may come from compact binary mergers).
Many additional mechanisms have been suggested over the years to produce γ-ray bursts,
and GRB 060614 may be the first clear example of one of these alternative scenarios.
It is surprising, however, that three distinct processes (collapsars, binary mergers, and
"the third mechanism") would results in events which are so similar in their high-energy
properties. A possible explanation would be that these are three different routes that
begin with very different physical systems but lead to a similar outcome, e.g., rapid
accretion onto a BH. The many similarities between short and long GRBs24 may hint at
this direction.
Regardless of the ultimate resolution of this puzzle (following studies of GRB 060614
and possible additional future examples of this class) it is already obvious that the elegant
simple picture, consisting of two groups of GRBs with distinct physical origins (long GRBs
from SN/collapsars and short GRBs from binary mergers) which was briefly consistent
with GRB observations and theory, must now be revised. ∗
Acknowledgements
A.G. acknowledges support by NASA through Hubble Fellowship grant #HST-HF-
01158.01-A awarded by STScI, which is operated by AURA, Inc., for NASA, under con-
tract NAS 5-26555. SRK's research is supported by NSF and NASA.
Received 11 June 2021; Accepted draft.
1. MacFadyen, A. I., Woosley, S. E. & Heger, A. Supernovae, Jets, and Collapsars. ApJ 550,
410 -- 425 March 2001.
2. Galama, T. J., Vreeswijk, P. M., van Paradijs, J., Kouveliotou, C., Augusteijn, T. et al. An
unusual supernova in the error box of the gamma-ray burst of 25 April 1998. Nature 395,
670 -- 672 October 1998.
∗Correspondence should be addressed to A. Gal-Yam ([email protected]).
6
Gal-Yam et al.
3. Stanek, K. Z., Matheson, T., Garnavich, P. M., Martini, P., Berlind, P. et al. Spectroscopic
Discovery of the Supernova 2003dh Associated with GRB 030329. ApJl 591, L17 -- L20 July
2003.
4. Hjorth, J., Sollerman, J., Møller, P., Fynbo, J. P. U., Woosley, S. E. et al. A very energetic
supernova associated with the γ-ray burst of 29 March 2003. Nature 423, 847 -- 850 June
2003.
5. Gehrels, N., Sarazin, C. L., O'Brien, P. T., Zhang, B., Barbier, L. et al. A short γ-ray burst
apparently associated with an elliptical galaxy at redshift z = 0.225. Nature 437, 851 -- 854
October 2005.
6. Bloom, J. S., Prochaska, J. X., Pooley, D., Blake, C. H., Foley, R. J. et al. Closing in on
a Short-Hard Burst Progenitor: Constraints from Early-Time Optical Imaging and Spec-
troscopy of a Possible Host Galaxy of GRB 050509b. ApJ 638, 354 -- 368 February 2006.
7. Berger, E., Price, P. A., Cenko, S. B., Gal-Yam, A., Soderberg, A. M. et al. The afterglow
and elliptical host galaxy of the short γ-ray burst GRB 050724. Nature 438, 988 -- 990
December 2005.
8. Parsons, A. M., Cummings, J. R., Gehrels, N., Goad, M. R., Gronwall, C. et al. GRB 060614:
Swift detection of a burst with a bright optical and X-ray counterpart. GRB Coordinates
Network 5252, 1 -- + (2006).
9. Golenetskii, S., Aptekar, R., Mazets, E., Pal'Shin, V., Frederiks, D. et al. Konus-wind
observation of GRB 060614. GRB Coordinates Network 5264, 1 -- + (2006).
10. Fox, D. B., Frail, D. A., Price, P. A., Kulkarni, S. R., Berger, E. et al. The afterglow of GRB
050709 and the nature of the short-hard γ-ray bursts. Nature 437, 845 -- 850 October 2005.
11. Mangano, V. e. a. Swift paper on GRB 060614, in preparation. To be submitted to ... , 1
(2006).
12. Pian, E., Mazzali, P. A., Masetti, N., Ferrero, P., Klose, S. et al. Gamma-Ray Burst asso-
ciated Supernovae: Outliers become Mainstream. Submitted to Nature, astro-ph/0603530 ,
.
13. Fynbo, J. P. U., Thoene, C. C., Jensen, B. L., Hjorth, J., Sollerman, J. et al. GRB060614:
detection of the host galaxy but no supernova emission. GRB Coordinates Network 5277,
1 -- + (2006).
14. Modjaz, M., Stanek, K. Z., Garnavich, P. M., Berlind, P., Blondin, S. et al. Early-Time
Photometry and Spectroscopy of the Fast Evolving SN 2006aj Associated with GRB 060218.
ApJL 645, L21 -- L24 July 2006.
15. Pastorello, A., Zampieri, L., Turatto, M., Cappellaro, E., Meikle, W. P. S. et al. Low-
luminosity Type II supernovae: spectroscopic and photometric evolution. MNRAS 347,
74 -- 94 January 2004.
A new explosive origin for GRBs
7
16. Christensen, L., Hjorth, J. & Gorosabel, J. UV star-formation rates of GRB host galaxies.
A&A 425, 913 -- 926 October 2004.
17. Gorosabel, J., P´erez-Ram´ırez, D., Sollerman, J., de Ugarte Postigo, A., Fynbo, J. P. U.
et al. The GRB 030329 host: a blue low metallicity subluminous galaxy with intense star
formation. A&A 444, 711 -- 721 December 2005.
18. Prochaska, J. X., Bloom, J. S., Chen, H.-W., Hurley, K. C., Melbourne, J. et al. The Host
Galaxy of GRB 031203: Implications of Its Low Metallicity, Low Redshift, and Starburst
Nature. ApJ 611, 200 -- 207 August 2004.
19. Zeh, A., Klose, S. & Hartmann, D. H. A Systematic Analysis of Supernova Light in Gamma-
Ray Burst Afterglows. ApJ 609, 952 -- 961 July 2004.
20. Soderberg, A. M., Kulkarni, S. R., Price, P. A., Fox, D. B., Berger, E. et al. An HST Study of
the Supernovae Accompanying GRB 040924 and GRB 041006. ApJ 636, 391 -- 399 January
2006.
21. Nomoto, K., Tominaga, N., Umeda, H., Maeda, K., Ohkubo, T. et al. Nucleosynthesis in
Hypernovae and Population III Supernovae. Nuclear Physics A 758, 263 -- 271 July 2005.
22. MacFadyen, A. I. in AIP Conf. Proc. 662: Gamma-Ray Burst and Afterglow Astronomy
2001: A Workshop Celebrating the First Year of the HETE Mission (eds Ricker, G. R. &
Vanderspek, R. K.) 202 -- 205 April 2003.
23. Narayan, R., Piran, T. & Kumar, P. Accretion Models of Gamma-Ray Bursts. ApJ 557,
949 -- 957 August 2001.
24. Nakar, E. Short/Hard Gamma-Ray Bursts. to be submitted to Physics Report , August 2006.
25. French, J., Melady, G., Hanlon, L., Jelinek, M. & Kubanek, P. GRB060614: watcher obser-
vation. GRB Coordinates Network 5257, 1 -- + (2006).
26. Gal-Yam, A., Leonard, D. C., Fox, D. B., Cenko, S. B., Soderberg, A. M. et al. On the
progenitor of SN 2005gl and the nature of Type IIn supernovae. Submitted to ApJ, astro-
ph/0608029 , .
8
Gal-Yam et al.
Figure 1: Temporal evolution of the optical transient associated with GRB 060614. The
main figure shows V-band observations from UVOT (solid circles), our SSO R-band data
(solid squares), augmented by by data from GCN 525725 (empty triangles) and GCN
527713 (empty diamonds), along with our late time HST V- and I-band detections (stars)
and upper limit (Solid inverted triangle; see Fig. 3). Note that our HST detections fall
below projections based on early V-band data or the fit proposed13 based on R-band data
(red dashed curve), indicating a steepening of the temporal decay prior to our first HST
visit. The contribution of the host galaxy, estimated as detailed in Fig. 2, was removed
from the photometry. The inset shows a comparison of our HST detections and upper
limit with scaled-down light curves of SNe, properly K-corrected and time-dilated.20 The
brightest allowed SN is obtained by assuming the steeply declining light curve of SN
1994I (heavy dash-dot) and the maximal amount of extinction allowed by the analysis
of early UVOT data (see supplementary Fig. 4). In this case the absolute magnitude of
the SN will be MV = −12.6, fainter than any SN ever detected in the nearby Universe,
synthesizing only ∼ 8 × 10−4 M⊙ of Ni 56. More likely scenarios involving moderate
extinction values (AV = 0.5) or SNe with more slowly-decaying light curves such as SN
2002ap (thin dashed line), similar to the faintest GRB-associated SN 2006aj, would impose
more stringent limits on the luminosity and Ni 56 production of a putative SN (by factors
of ∼ 3). Note that in all cases the emission detected during our first HST visit must be
dominated by the GRB afterglow emission, with but a fraction of the light coming from
the peaking faint SN, and that the data are well-explained without invoking any SN-like
component, with the optical afterglow roughly following the late X-ray decline rate11 (to
which a SN is not expected to contribute).
A new explosive origin for GRBs
9
−15
−16
−17
−18.5
−18
−19
−19.5
−19
−20
−20
SN 1994I / 100, I
SN 2002ap / 132, I
SN 1994I / 100, V
SN 2002ap / 132, V
/
g
n
A
2
m
c
/
c
e
s
/
g
r
e
n
i
λ
F
g
o
l
−20.5
5
−21
2
2.5
3
Figure 1.
5.5
6
6.5
3.5
log time since GRB in s
4.5
5
4
Swift V
SSO R
Watcher R
Danish R
HST I
HST V
5.5
6
6.5
10
Gal-Yam et al.
Figure 2: A spectrum of the host galaxy of GRB 060614 (thin blue curve) obtained with
the GMOS-S spectrograph mounted on the Gemini-South 8m telescope at Cerro Pachon,
Chile, on July 15, 2006 UT. Four exposures of 1200s each were reduced and combined
in the usual manner within IRAF, including wavelength- and flux-calibration using the
smooth spectrum standard star EG131. The spectral continuum is similar to a template
Sc galaxy spectrum (heavy green curve) shifted to z = 0.125 and scaled in flux. We note
that the emission lines are much weaker, though. From the luminosity of the Hα line we
derive an upper limit on the star-formation rate in the galaxy, SFR = 0.0035 M⊙ yr−1.
The ratio of Hα to Hβ line strengths indicates that the emission-line regions in this dwarf
galaxy suffer negligible extinction. Assuming that the galaxy spectrum is well-described
by the Sc template throughout the optical range, we use synthetic photometry anchored
to the V-band magnitude of the galaxy measured from our HST images to determine the
magnitudes of the host to be [UBVRI]=[23.24 23.89 23.66 23.47 22.87] mag, respectively.
The host contribution is removed from the photometry presented in Fig. 1.
A new explosive origin for GRBs
11
x 10−18
12
]
/
g
n
A
2
m
c
/
s
/
g
r
e
[
λ
F
10
8
6
4
2
4500
5000
Figure 2.
5500
6000
Observed Wavelength [Ang]
6500
7000
7500
8000
12
Gal-Yam et al.
Figure 3.
Figure 3: HST observations of the location of GRB 060614. Upper panels show I-
band (F814W filter) data and lower panels V-band (F606W) data. Images were obtained
using the WFPC2 camera (left, 6000s total exposure time; F814W filter; UT dates as
marked) and the ACS camera (middle panels; total exposure 3600s; same filter). The
third panels show the difference between the first epoch images and a third epoch visit
with ACS (I-band; July 29, 2006 UT, total exposure time 4840s) or the second epoch
in the case of the V-band, calculated using our image subtraction methods.26 WFPC2
data were reduced and photometered using an adaptation of HSTphot26 and ACS data
were reduced in the standard manner with IRAF/multidrizzle. We have used HSTphot-
calibrated, nearby, isolated, compact sources to establish a calibration grid of Johnson-V
and Cousins-I local standards and photometered the afterglow with respect to this grid
using the image-subtraction-based photometry pipeline mkdifflc.26 A similar comparison
between the second and third I-band HST visits (July 16 and 29 UT, respectively) shows
no residual to a 4σ upper limit of I=28.1 mag, indicating that the optical transient was
undetectable during our second visit and justifying the use of the second-epoch V-band
image as a subtraction template. The resulting photometry and upper limit are reported
in Fig. 1. Note the overall regular structure of this faint dwarf host and the peripheral
location of the optical transient. The orientation of the images is marked with a long
arrow due north and a short arrow due east. The length of the long arrow is 2.5′′ for
scale.
A new explosive origin for GRBs
13
UVOT photometric spectum, AV=0.5
ν0.5
UVOT photometric spectum, AV=1.7
ν1.94
−25.4
−25.6
−25.8
−26
−26.2
−26.4
−26.6
/
z
H
2
m
c
/
c
e
s
/
g
r
e
n
i
ν
F
g
o
l
−26.8
−27
−27.2
14.7
14.8
14.9
15
log frequency [Hz]
15.1
15.2
15.3
Figure 4.
Supplementary Figure 4: A photometric spectrum derived from early multi-band
UVOT observations. The flux in the 5 broad-band UVOT filters (V,B,U,UVW1,UVW2
with effective wavelengths of 550, 450, 330, 250 and 180 nm, respectively) was interpolated
to 10000s after the GRB. We consider a hot thermal spectrum (Fν ∝ ν2) to be the bluest
possible intrinsic spectrum at this time (such a blue spectrum has never been observed in
any GRB). The observed spectrum cannot be dereddened by more than AV = 1.7 without
becoming bluer than this hot spectrum (red curve and tiangles). A more typical blue
intrinsic spectrum for a GRB afterglow with Fν ∝ ν0.5 allows only moderate extinction
(AV = 0.5). Fits of similar quality are also obtained without invoking any extinction. At
later times (up to 105 s) the observed photometric spectrum becomes redder, imposing
weaker constraints on the extinction. We adopt AV = 1.7 as a conservative upper limit
on the extinction toward this optical transient, but estimate that the more likely value is
actually lower (AV ≈ 0.5 or less).
This figure "fighst.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0608257v1
|
astro-ph/0403587 | 1 | 0403 | 2004-03-25T05:27:48 | Interpretation of parabolic arcs in pulsar secondary spectra | [
"astro-ph"
] | Pulsar dynamic spectra sometimes show organised interference patterns; these patterns have been shown to have power spectra which often take the form of parabolic arcs, or sequences of inverted parabolic arclets whose apexes themselves follow a parabolic locus. Here we consider the interpretation of these arc and arclet features. We give a statistical formulation for the appearance of the power spectra, based on the stationary phase approximation to the Fresnel-Kirchoff integral. We present a simple analytic result for the power-spectrum expected in the case of highly elongated images, and a single-integral analytic formulation appropriate to the case of axisymmetric images. Our results are illustrated in both the ensemble-average and snapshot regimes. Highly anisotropic scattering appears to be an important ingredient in the formation of the observed arclets. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000–000 (0000)
Printed 14 June 2018
(MN LATEX style file v2.2)
Interpretation of parabolic arcs in pulsar secondary spectra
M.A. Walker1, D.B. Melrose1, D.R. Stinebring2, C.M. Zhang1
1. School of Physics, University of Sydney, NSW 2006, Australia
2. Oberlin College, Department of Physics and Astronomy, Oberlin, OH 44074, U.S.A.
14 June 2018
ABSTRACT
Pulsar dynamic spectra sometimes show organised interference patterns; these patterns have
been shown to have power spectra which often take the form of parabolic arcs, or sequences of
inverted parabolic arclets whose apexes themselves follow a parabolic locus. Here we consider
the interpretation of these arc and arclet features. We give a statistical formulation for the
appearance of the power spectra, based on the stationary phase approximation to the Fresnel-
Kirchoff integral. We present a simple analytic result for the power-spectrum expected in the
case of highly elongated images, and a single-integral analytic formulation appropriate to the
case of axisymmetric images. Our results are illustrated in both the ensemble-average and
snapshot regimes. Highly anisotropic scattering appears to be an important ingredient in the
formation of the observed arclets.
Key words: Scintillations - pulsars - interstellar medium
1 INTRODUCTION
At low radio frequencies, diffraction from inhomogeneities in the
ionised component of the InterStellar Medium (ISM) leads to
multi-path propagation from source to observer. Together with the
refraction introduced by large-scale inhomogeneities, this leads to
a variety of observable phenomena - image broadening, image
wander and flux variations, for example, see Rickett (1990). If the
source is very small, as in the case of radio pulsars, then the elec-
tric field has a high level of coherence amongst the various propa-
gation paths, and persistent, highly visible interference fringes are
seen in the low frequency radio spectrum. These fringes evolve
with time, due to the motions of source, observer and ISM. Mostly
the observed fringes can be understood in terms of wave propa-
gation through a random medium, with the electron density inho-
mogeneities being described by a power-law spectrum, close to the
Kolmogorov spectrum, and physically identified with a turbulent
cascade (Armstrong, Rickett and Spangler 1995; Cordes, Weisberg
and Boriakoff 1985; Lee and Jokipii 1976). However, it has been
known for many years that some pulsars, at some epochs, also ex-
hibit organised patterns in their dynamic spectra – drifting bands,
criss-cross patterns and periodic fringes, for example (Gupta, Rick-
ett and Lyne 1994; Wolsczan and Cordes 1987; Hewish, Wol-
szczan and Graham 1985; Roberts and Ables 1982; Ewing et al
1970). The interpretation of these patterns has never been entirely
clear, although several authors have suggested an association with
enhanced refraction in the ISM (Rickett, Lyne and Gupta 1997;
Hewish 1980; Shishov 1974).
Recently a significant breakthrough in data analysis has oc-
curred, with the recognition that these organised patterns are more
readily apprehended by working with the so-called "secondary
spectrum", i.e. the power spectrum of the dynamic spectrum (Stine-
bring et al 2001): in this domain the very complex patterns seen
in the dynamic spectrum typically appear as an excess of power
along the locus of either a single parabola, or a collection of in-
verted parabolic arcs (Hill et al 2003; Stinebring et al 2004). Pre-
vious investigations have demonstrated why parabolic structures
are generic features of the secondary spectrum (Stinebring et al
2001; Harmon and Coles 1983), but the properties of the observed
secondary spectra are not yet understood in detail. In this paper
we develop a statistical theory of the power distribution in pulsar
secondary spectra, based on the stationary phase approximation,
and present some analytic results for simple forms of the scattered
image. We compare our results with existing data and argue that
highly anisotropic scattering is required to form some of the ob-
served secondary spectra. Our results are consistent with, and com-
plementary to, those of Cordes et al (2003).
This paper is organised as follows. We begin by developing
a model secondary spectrum, based on the stationary phase ap-
proximation, in §2, leading to analytic models for the ensemble-
average secondary spectra of linear and axisymmetric images (§3),
and Monte Carlo simulations of snapshot secondary spectra (§4).
In §5 we consider the appearance of additional, off-centre image
components; we compare our models with existing data in §6.
2 THE STATIONARY PHASE APPROXIMATION
Consider a radio wave propagating over a distance Dp from a
source to the observer, and passing en route through a phase-
changing screen at a distance Ds from the observer. We consider
only the case of a point-like source, implying a large coherent
patch; this is a sensible approximation for pulsars, which are known
to be extremely compact sources. Let x denote the 2D position on
the phase screen, then the wave amplitude at a point r in the ob-
2 Walker et al.
server's plane is
u(r) = −i
2πr2
F Z d2x exp(iΦ),
and to high accuracy the phase, Φ, is given by
(x − βr)2
Φ = φ(x) +
,
2r2
F
(1)
(2)
with r2
F := β Ds/k , β ≡ 1 − Ds/Dp, and k = 2πν/c for fre-
quency ν. Here the pulsar is assumed to lie at the origin of the trans-
verse coordinates (r and x). In the trivial case where φ = 0 every-
where this yields unit wave amplitude at the observer - the usual
spherical wave amplitude variation, exp(i kz)/z, having been ab-
sorbed into this normalisation.
For many purposes the exact integral formulation of eq. 1 can
be adquately represented by a sum over a finite number of discrete
points – the stationary phase points – for which the derivative of
the exponent with respect to the variable of integration vanishes
(see, e.g., Gwinn et al 1998). This replacement can be made be-
cause these points, and their immediate surroundings, dominate the
integral: where the phase is not stationary, the integrand oscillates,
changing sign rapidly, and there is little net contribution to the in-
tegral. The condition for a point of stationary phase is ~∇Φ = 0, or
~∇φ +
x − βr
r2
F
= 0.
(3)
For each solution of equation 3, x = xi(r) with i = 1...N, we ex-
pand the integrand to second order in x about the stationary point. It
is assumed that these points are all well separated, so that this pro-
cedure may be applied to each point separately. The integral taken
around the i th point gives a contribution
ui(r) = √µi exp(iΦi),
(4)
where µi is the "magnification", and Φi the phase for each path.
The magnification is determined from the phase curvature intro-
duced by the screen:
µ−1 = (1 − κ)2 − γ2,
(5)
in terms of the convergence, κ, and shear γ :=pγ2
1 + γ2
2 , with
κ = −
r2
F
2 (cid:20) ∂2φ
∂x2 +
∂2φ
∂y2(cid:21)
and
γ1 = −
r2
F
2 (cid:20) ∂2φ
∂x2 −
∂2φ
∂y2(cid:21) ,
γ2 = −r2
F
∂2φ
∂x∂y
,
where x = (x, y). The total electric field is then
(6)
(7)
(8)
√µiµj cos Φij ,
(9)
and the total intensity is
I(ν, t) = u∗u ≃(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
u(r) =Xi
ui(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi=1
N
2
ui(r),
N
=
Xi,j=1
with Φij ≡ Φi − Φj. This result is valid not only for the case
of strong scattering, where N ≫ 1 – with each of the N points
coinciding with a speckle in the image – but also for lens-like (re-
fractive) behaviour where there may be only a very small number
of stationary phase points.
The result we have just derived is a description of the dynamic
spectrum of the source, because each of the Φi is a function of
frequency and time. We will assume that the magnifications change
only slowly in comparison with the phases. In some circumstances
– when the point under consideration is close to a critical curve
(corresponding to the source lying close to a caustic) – the µi are
expected to change rapidly with time and frequency; we will not
consider such cases, and henceforth we take the µi to be constant.
We can approximate the Φ variations as linear in frequency and
time, over a total observing time ∆t, centred on t0, and bandwidth
∆ν centred on ν0:
ij +
Φij ≃ Φ0
ij ≡ Φij (ν0, t0). Taking the Fourier Transform of equa-
(t − t0) +
(ν − ν0),
(10)
where Φ0
tion 9 then yields I := I(fν, ft):
∂Φij
∂ν
∂Φij
∂t
∆t ∆ν
I =
N
Xi,j=1
√µiµj(cid:8)exp[iΦ0
ij ] sinc[π ∆t (ft + Ft,ij )]
4π
× sinc[π ∆ν (fν + Fν,ij )]
× sinc[π ∆t (ft − Ft,ij)] sinc[π ∆ν (fν − Fν,ij )](cid:9). (11)
For large ∆t, ∆ν this result can be approximated by
exp[−iΦ0
ij ]
+
ij ] δ(ft + Ft,ij ) δ(fν + Fν,ij )
I ≃
1
4π
N
Xi,j=1
√µiµj(cid:8)exp[iΦ0
+ exp[−iΦ0
where δ denotes the Dirac Delta Function, and
ij ] δ(ft − Ft,ij ) δ(fν − Fν,ij )(cid:9), (12)
Ft,ij ≡
1
2π
∂Φij
∂t
,
Fν,ij ≡
1
2π
∂Φij
∂ν
.
(13)
Our model "secondary spectrum" is therefore (the power spectrum
of equation 9):
N
∆t ∆ν
(4π)2
(14)
P (fν, ft) = I ∗ I ≃
µiµj(cid:2)δ (ft + Ft,ij )
×δ (fν + Fν,ij ) + δ (ft − Ft,ij ) δ (fν − Fν,ij )(cid:3).
Xi,j=1
The foregoing analysis shows that the various interference
terms contribute power to the secondary spectrum in proportion
to µiµj, at specific fringe frequencies given by the derivatives of
the phase differences, Φij. We note that the secondary spectrum is
symmetric under the operation (fν, ft) → (−fν,−ft). This arises
because I(ν, t) is a real quantity, reflecting the interchange asym-
metry Φij = −Φji. Another manifestation of the symmetry is the
fact that fringes with wavevector (fν, ft), and those with wavevec-
tor (−fν, −ft) are indistinguishable in the dynamic spectrum. Be-
cause of the symmetry in P , we will henceforth consider only the
half-plane fν > 0, for which the secondary spectrum just derived
can be rewritten as
P (fν, ft) ∝
µiµj δ (ft − Ft,ij) δ (fν − Fν,ij ) ,
(15)
and this is the form we shall make use of subsequently.
N
Xi,j=1
2.1 Phase relationships in multipath propagation
The trajectory associated with the i th stationary phase point ex-
hibits a phase Φi = Φ(xi, r, ν, t), given by equation 2, relative to
the direct line-of-sight to the source in the absence of a screen. The
derivatives of Φ with respect to these four variables can be eval-
uated as follows. Quite generally, differentiating equation 2 with
Parabolic arcs
3
frequency shifts have not been calculated from a relativistic formu-
lation, and are therefore not correct to second order in v/c.
(16)
The spectral fringe rate for stationary phase paths,
respect to r yields
∂r(cid:19)x,ν,t
(cid:18) ∂Φ
= −
β
r2
F
(x − βr).
Similarly, differentiating equation 2 with respect to ν gives
∂ν(cid:19)x,t,r
(cid:18) ∂Φ
=
1
ν (cid:20) (x − βr)2
2r2
F
where we have made use of the result
(cid:18) ∂ log φ
∂ log ν(cid:19)t
= −2,
− 2φ(cid:21) ,
(17)
(18)
which is appropriate for radio-wave propagation in the Galaxy
where the refractive index is dominated by the free-electron con-
tribution. Differentiating equation 2 with respect to x gives
(cid:18) ∂Φ
∂x(cid:19)ν,t,r
= ~∇φ +
x − βr
r2
F
= 0,
(19)
where the final equality (equation 3) applies for stationary phase
points. Finally, differentiating equation 2 with respect to t yields
∂t (cid:19)x,r,ν
(cid:18) ∂Φ
=(cid:18) ∂φ
∂t(cid:19)x,ν
.
(20)
We will assume that the phase structure in the screen is "frozen",
and thus is convected with the screen velocity, v′
s, so that
s · ~∇φ +
v′
∂φ
∂t
= 0
(21)
for all paths. Using the stationary phase condition then gives us
(cid:18) ∂Φi
∂t (cid:19)x,r,ν
=
1
r2
F
(x − βr) · v′
s.
Making use of equations 16 and 22, it is a straightforward ex-
ercise to evaluate the temporal fringe frequency for a given station-
ary phase path:
=
+ v′
1
r2
F
=(cid:18) ∂Φi
∂t (cid:19)x,r,ν
∂r(cid:19)x,ν,t
s − βv′
o),
∂t (cid:19)ν
2π Ft,i =(cid:18) ∂Φi
o ·(cid:18) ∂Φ
(x − βr) · (v′
where the observer's velocity is v′
o. If the source velocity is vp, rel-
ative to some reference frame, then the actual space velocities of the
screen and observer are vs,o = v′
s,o + vp. Only the components of
these velocities perpendicular to the line-of-sight are of relevance
here. Defining v⊥ ≡ vs − βvo − (1 − β)vp, where all velocities
are to be understood as being transverse components, allows us to
write (dropping the subscript i which denotes the stationary phase
path under consideration)
(23)
Ft =
1
λβ
θ · v⊥,
(24)
where λ = 2π/k, and θ = (x − βr)/Ds is the angular separation
between the path under consideration and the direct line-of-sight to
the source.
Implicit in this result for the temporal fringe frequency are the
Doppler shifts due to the motions of source, screen and observer,
each of which contributes to v⊥. All of these terms are linear in
the relevant velocity and in the angular separation of the paths un-
der consideration, so for speeds ∼ 300 km s−1, typical of pulsars,
and angles of order a couple of milli-arcseconds, the fractional fre-
quency shifts are ∼ θv/c ∼ 10−11, hence beat periods of order
two minutes at a radio frequency of 1 GHz. We note that these
(22)
2.2 Neglect of dispersive delays
Fν,i ≡
1
2π (cid:18) ∂Φi
∂ν (cid:19)t
=
1
2π (cid:18) ∂Φ
∂ν(cid:19)x,t,r
,
(25)
follows directly from equation 17. Again dropping the subscript
which denotes the particular stationary phase path under consider-
ation, we have
Fν =
Dsθ2
2cβ −
φ
πν
.
(26)
The physical interpretation of the various terms contributing to Fν
is straightforward. The first term in equation 26 is the geometric
contribution to the delay along the path under consideration, while
the second term is the delay due to the propagation speed of the
radio-wave within the phase screen; the former is frequency inde-
pendent, while the latter is dispersive. The geometric delay is ex-
pected to be ≃ θ2D/c ∼ 10−5s, for refraction angles θ ∼ 2 milli-
arcseconds, and source distance ∼ 1 kpc.
Equations 24 and 26 apply to the individual stationary phase
paths, from which we can compute Ft,ij, Fν,ij, via Φij ≡ Φi−Φj.
Having identified the physical interpretation of the terms contribut-
ing to Ft, Fν, as being Doppler shifts and delays, respectively,
we see that the secondary spectrum can be regarded as a "delay-
Doppler diagram", with the power appearing at locations corre-
sponding to differences in delay/Doppler-shift between the various
interfering paths (Harmon and Coles 1983; Cordes et al 2004).
From equations 24 and 26 it can be seen that if the dispersive delay
(the last term in eq. 26) is neglected, then the fringe frequencies
obey a quadratic relationship because Ft depends linearly on θ,
whereas Fν varies quadratically. Neglect of the dispersive delay is
a good approximation, for stationary phase paths, if
(x − βr) ·
≫ 4.
(27)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
~∇φ
φ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
In other words the screen phase, φ, must be changing on a length-
scale (L) which is small compared with the distance of the path
from the optical axis (the direct line-of-sight to the source). The
meaning of this result can be seen most easily if we consider two
limiting cases: purely scattering (diffractive), and purely lens-like
(refractive) phase screens. In the pure scattering case, ~∇φ changes
sign on a length scale ∼ L ≪ rF much smaller than the typical
separation x − βr & rF . Consequently the neglect of dispersive
delays is a natural approximation to make in the case of a pure
scattering screen. However, the converse is not necessarily true; to
see this we can consider a screen in which the phase increases as
(x − βr)n. In this case the condition given in equation 27 becomes
n ≫ 4, which clearly can be satisfied even though there is no small-
scale phase structure (no scattering) in the phase screen.
In the major part of the present paper we will confine our-
selves to consideration of the case where the dispersive delays are
negligible, returning to the issue only in §5. This restriction is mo-
tivated in large part by the attendant simplification of results, and
encouraged by the fact that the resulting model seems to offer a
fair description of many of the observed phenomena. Noting that
the fringe frequencies Ft,ij and Fν,ij include common normalis-
ing factors (eqs 24 and 26), it is useful to introduce the following
4 Walker et al.
coordinates (Stinebring et al 2001; Hill et al 2003):
and
q :=
βλ
v⊥
ft = θ1x − θ2x,
p :=
2cβ
Ds
fν = θ2
1 − θ2
2.
(28)
(29)
For simplicity we have chosen the x-axis to lie along v⊥.
3 STATISTICAL MODEL OF THE SECONDARY
SPECTRUM
Distant sources (D & 1 kpc) observed at low radio frequencies are
expected to be in the regime of strong scattering, for which there
are many paths of stationary phase from source to observer. In this
circumstance it is appropriate to employ a statistical treatment of
the power distribution in the secondary spectrum. In this section
we present such a treatment, with emphasis on the analytic results
which can be obtained for some specific cases. These results should
prove to be useful for observers undertaking quantitative analysis
of the power distribution in secondary spectra.
The intensity distribution in the scattered image can be writ-
ten in terms of a probability distribution, g(θ), for the power in
the stationary phase paths. Equation 15 shows that the various in-
terference terms contribute power to the secondary spectrum in
proportion to µiµj, and the expectation value of this quantity is
g(θ1)g(θ2) d2θ1d2θ2, where we have replaced the specific labels
i, j by the generic pair label 1, 2. Including only interference terms
which contribute to the particular fringe frequencies p, q consid-
ered, the expected power distribution in the secondary spectrum
can therefore be described by
Figure 1. "Secondary spectrum" (or "delay-Doppler" diagram) for a linear
image, at an angle ψ = 0 to the x-axis (defined by the effective transverse
velocity vector), with a Gaussian intensity profile (defined by equation 39).
The axes are expressed in units of θo = 1/k so, for q, and θ2
o for p. Contour
levels are set at intervals of 3 dB, and the lowest (outermost) contour is at
−60 dB relative to P (0.1, 0.1).
data. However, when comparing our models with data we must bear
in mind that some deviations are expected simply because the data
are not ensemble averages.
3.1 Image-plane probability distributions
In order to compute the secondary spectrum via equation 30, we
need to specify the probability distribution g(θ). The expected an-
gular distribution of power is given by the spatial Fourier Transform
of the mutual coherence function (Lee and Jokipii 1975):
1 + θ2
P =Z d2θ1d2θ2 g(θ1)g(θ2) δ(p − θ2
2)δ(q − θ1x + θ2x).
(30)
Strictly this formulation should be interpreted as yielding the
ensemble-average (i.e. long-term time-average) of P . However, in
order to reach this regime observationally one must average data
over time-scales which are long in comparison to the refractive
time-scale, and this is generally not the case in practice. In prin-
ciple one could average the secondary spectra for a given target
observed over a period of several months or years, much longer
than the refractive time-scale, but in practice this is not a sensible
procedure because the large-scale distribution of power in the sec-
ondary spectrum can undergo profound changes on time-scales of
weeks, indicating that the statistical properties are not stationary
on these time-scales, so that the ensemble-average regime cannot
be reached. Instead we are obliged to deal with data in the limit of
almost instantaneous sampling, or else in the average-image regime
with only a short averaging time. The various regimes are charac-
terised by an averaging time, t, which stands in the following rela-
tionships to the diffractive time-scale, td, and the refractive time-
scale, tr (Goodman and Narayan 1989): t ≫ tr, ensemble aver-
age; tr & t > td, average; t < td, snapshot. For pulsars observed
at frequencies ν ∼ 400 − 1000 MHz, representative time-scales
are td ∼ minutes, tr ∼ weeks; a single observation might last for
t ∼ 30 minutes, and is therefore generally in the average-image
regime. Some data may be in the snapshot regime, and this regime
is considered in §4. The expected image-plane power distributions
in the average and ensemble-average regimes are identical, and thus
the formulation given in equation 30 is appropriate for much of the
g(θ) ∝Z d2ρ exp(−ikρ · θ) Γ2(ρ).
(31)
For scattering by turbulence in a single screen, the mutual coher-
ence function is
Γ2(ρ) = exp(−D(ρ)),
where the phase structure function is given by
D(ρ) = h[φ(x + ρ) − φ(x)]2i.
(32)
(33)
In the present paper we will consider only structure functions which
are power-law in the separation ρ, and we parameterise these forms
in terms of the field coherence scale, so: D(so) = 1. A character-
istic angular scale for the image is then θo = 1/(k so).
3.2 Highly anisotropic images
A useful model for highly anisotropic images follows when one
considers the limiting case of a linear image. A linear source mak-
ing an angle ψ with respect to the x-axis corresponds to an image
with
g(θ) = h(θx cos ψ + θy sin ψ) δ(−θx sin ψ + θy cos ψ),
(34)
where the function h denotes the variation in intensity along the
line. On inserting eq. 34 into eq. 30, there are four integrals and
four δ-functions. The probability P (p, q) is then found by solving
the four simultaneous equations, implied by the four δ-functions,
for θ± = θx± cos ψ + θy± sin ψ and identifying P (p, q) as
h(θ+)h(θ−) divided by the Jacobian of the transformation to the
Parabolic arcs
5
Figure 2. Secondary spectrum for a linear image, as figure 1 but for the
image intensity profile given by equation 41 (representing the case of scat-
tering by highly anisotropic Kolmogorov turbulence). Note the difference
in the scale of the axes relative to figure 1; the contour levels are identical
to those in figure 1.
Figure 3. "Secondary spectrum" (or "delay-Doppler" diagram) for an ax-
isymmetric image with a Gaussian intensity profile (defined by equation
47). The axes are expressed in units of θo = 1/kso, for q, and θ2
o for
p. Contour levels are set at intervals of 3 dB, and the lowest (outermost)
contour is at −60 dB relative to P (0.1, 0.1).
variables specified by the δ-functions. The four simultaneous equa-
tions give
p cos2 ψ + q2
p cos2 ψ − q2
2q
2q
,
,
θy+ = tan ψ θx+,
θy− = tan ψ θx−,
θx+ =
θx− =
and hence
θ± =
p cos2 ψ ± q2
2q cos ψ
.
The Jacobian is q, hence one finds
P (p, q) =
with θ± given by eq. 36.
h(θ+) h(θ−)
q
,
(35)
(36)
(37)
3.2.1 Specific examples of anisotropic images
Here we take the general formulation given above and apply it to
two specific cases. For simplicity we calculate only the case ψ =
0, corresponding to an image which is elongated along the same
direction as the effective velocity v⊥.
In the case of a quadratic phase structure function,
so(cid:19)2
D(ρ) =(cid:18) ρx
,
(38)
with no dependence on ρy, the resulting image is Gaussian, with
h(θ) =
1
√2πθo
exp(−θ2/4θ2
o ).
(39)
The secondary spectrum corresponding to this image profile is
shown in figure 1, with q measured in units of θo, and p in units
of θ2
o. The symmetry of the secondary spectrum is due to the sym-
metry of the image around θx = 0. The contours in figure 1 are
spaced at intervals of 3 dB, with the lowest (outermost) contour
corresponding to −60 dB relative to the peak - the features shown
here are very faint. In this, and subsequent figures, we have taken
the peak value of the secondary spectrum to be P (0.1, 0.1); nor-
malising to these coordinates is somewhat arbitrary, but normalis-
ing at (0, 0) is not sensible because P → ∞ there. The fact that P
Figure 4. Secondary spectrum for an axisymmetric image, as figure 3 but
for the image intensity profile given by equation 48 (representing the case
of scattering by isotropic Kolmogorov turbulence). Note the difference in
the scale of the axes relative to figure 3; the contour levels are identical to
those in figure 3.
peaks for small p, q simply reflects the fact that the sum of the self-
interference terms is larger than the sum of cross-power terms in
equation 15. The most striking feature of figure 1 is the absence of
power in the region q = 0, p > 0; this is a consequence of the fact
that we are considering a linear image. The case q = 0 corresponds
to interference between points which have the same value of θx, but
we are considering an image where all of the power is concentrated
along a line with θy ∝ θx, so the secondary spectrum is devoid of
cross-power at small q - there is only the self-interference which
appears at p = 0.
A particularly interesting case to consider is that where the
structure in the phase screen arises from highly anisotropic Kol-
mogorov turbulence, in which case
D(ρ) =(cid:18) ρx
so(cid:19)5/3
(40)
(Narayan and Hubbard 1988), and the image profile cannot be ex-
pressed in a simple analytic form. Numerically we find that in this
case the intensity profile assumes a power-law form at large angles,
6 Walker et al.
with a power-law index of −8/3, and we adopt the approximate
form
h(θ) ≃
h(0)
1 + (θ/θo)8/3 ,
(41)
where h(0) ≡ 4 sin(3π/8)/3π. This image profile yields much
more power scattered to large angles than the Gaussian profile con-
sidered above. The corresponding secondary spectrum is displayed
in figure 2; note the large increase in the area plotted, relative to
figure 1, even though the contour levels are identical in the two
figures. This feature is a direct consequence of the fact that the
Kolmogorov phase screen yields an image with much more power
at large angles than the Gaussian profile, yielding much larger de-
lays (i.e. much larger p) at a given contour level. Figure 2 displays
the same absence of power in the region p > 0, q = 0 as seen in
figure 1, and for the same reasons. A new feature in figure 2, how-
ever, is the parabolic form p = q2 delineated by the contouring at
p, q ≫ 1. This can be understood by recognising that the dominant
contributions to secondary spectrum are due to interference terms
which include the highest intensity region of the image, namely
θ ≪ θo in this case. Referring to equation 36, we see that this
corresponds to θ− ≪ θo, hence p − q2 . 2q. More generally,
for a linear image making an angle ψ with the x-axis, at large val-
ues of p the secondary spectrum will exhibit contours which cluster
around p = q2 sec2 ψ.
3.3 Axisymmetric images
A priori perhaps the most plausible model is one in which the prob-
ability does not depend on direction: g(θ) = g(θ), i.e. the axisym-
metric case. In this case two of the integrals in eq. 30 can be per-
formed over the δ-functions, and a useful result requires that one
perform a third integral analytically. This would then leave a sin-
gle integral that can be evaluated either numerically, or with the
use of simple analytic approximations. In view of the symmetry,
the choice of polar coordinates is the most appropriate. With this
choice eq. 30 becomes
P (p, q) =Z dθ1θ1dψ1Z dθ2θ2dψ2 g(θ1) g(θ2)
× δ(p − θ2
2) δ(q − θ1 cos ψ1 + θ2 cos ψ2).
The integrals are carried out in the Appendix; the result is
1 + θ2
(42)
P (p, q) = 2Z Θ
0
+Z ∞
(p−q2)/2q
dθ2 θ2 g(θ2)g(cid:0)pp + θ2
2(cid:1)
[(pp + θ2
2 + θ2)2 − q2]1/2
×K
2 + θ2)2 − q2#1/2
"
4pp + θ2
2 θ2
(pp + θ2
dθ2 θ2 g(θ2)g(cid:0)pp + θ2
2(cid:1)
[pp + θ2
2 θ2]1/2
×K
" (pp + θ2
2 + θ2)2 − q2
4pp + θ2
#1/2
2 θ2
, (43)
where K is the Complete Elliptic Integral, and the upper limit of
integration is Θ = (p − q2)/4q + (p − q2)/4q. This result is
not amenable to further analytic manipulation, in general, and is
our final form for general axisymmetric images. In some cases it is
possible to approximate equation 43 with a simpler analytic result.
For example: if the probability distribution, g(θ), is very sharply
peaked at θ = 0, then for p, q > 0 and p > q2 the second integral
in eq. 43 can be neglected (see equations A.11 and A.12), and we
can make the approximation K ≃ π/2, so that
P (p, q) ≃ 2Z ∞
=
dθ2 θ2 g(θ2) g(√p)
π
2
0
pp − q2
1
2
g(√p)
pp − q2
.
(44)
However, in general the model secondary spectrum for any given
axisymmetric image must be evaluated numerically from the for-
mulation given in equation 43
3.3.1 Specific examples of axisymmetric images
For the general case of an isotropic, power-law structure function:
D(ρ) =(cid:18) ρ
so(cid:19)α
,
we have
g(θ) ∝Z d2ρ exp[−ikρ · θ − D(ρ)],
dt t J0(θt/θo) exp(−tα).
∝Z ∞
0
(45)
(46)
For α = 2 it is straightforward to show that the resulting image
profile is Gaussian, with
g(θ) =
1
4πθ2
o
exp(−θ2/4θ2
o).
(47)
The secondary spectrum corresponding to this image is shown in
figure 3; comparing this to figure 1 (the result for a linear, Gaussian
image) we see that the structure is broadly similar in the region
p . q2, but the axisymmetric image exhibits much more power in
the region p > q2. This can be understood in the following way:
whereas the linear images considered in §3.2.1 possess a unique
relationship of the form θy ∝ θx, so that a small value of q neces-
sarily corresponds to a small value of p, no such relationship exists
for the axisymmetric image - even for q = 0 (so θx1 = θx2),
there are many contributing values of θy1, θy2, and hence there is
power spread over a range of values of p.
By contrast to the case α = 2, for isotropic Kolmogorov tur-
bulence, with α = 5/3, the integral does not yield a simple ana-
lytic form. As with the highly anisotropic case in §3.2.1, evaluat-
ing numerically reveals a power-law variation in the image surface-
brightness, with g(θ) ∝ θ−11/3 for θ ≫ θo - there is much more
power scattered to large angles in the case of a Kolmogorov spec-
trum of turbulence than for the case of a quadratic phase structure
function. We will use the following approximation for the case of
isotropic Kolmogorov turbulence:
g(θ) ≃
g(0)
1 + (θ/θo)11/3 ,
(48)
where g(0) ≡ 11 sin(6π/11)/6π2. The corresponding secondary
spectrum is shown in figure 4; note the much expanded scale on
both axes, relative to figure 3, and the enhanced power around the
parabolic arc defined by p = q2. Comparing to the case of highly
anisotropic Kolmogorov turbulence, shown in figure 2, we find that
the axisymmetric case is similar to the anisotropic case for p . q2,
but exhibits much more power at p > q2. This same feature was
found in the comparison between linear and axisymmetric Gaus-
sian images (see above), and occurs for the same reason.
4 SNAPSHOT SECONDARY SPECTRA
The results described in §3 are expectation values for the secondary
spectrum in various circumstances; they represent the ensemble av-
erage – i.e. the average over many realisations of individual phase
Parabolic arcs
7
4.1 Monte Carlo realisations of snapshot spectra
To demonstrate the appearance of a secondary spectrum from a
single realisation of a phase screen, rather than the long-term av-
erage value, we return to the formulation given in §2. To generate
a secondary spectrum from equation 15 we need to specify the lo-
cations and magnifications for each of the stationary phase points
contributing to the image. If the expected image profile is specified,
e.g. a Gaussian profile in the case of a quadratic phase structure
function, then we can generate an individual realisation, consistent
with the expected properties, by the Monte Carlo method. The ex-
pected magnification of a stationary phase point is independent of
its location for the models we are employing in this paper, and for
simplicity we have adopted unit magnification for all the station-
ary phase points. The coordinates of the stationary phase points
can be generated as Normally-distributed random numbers, using
a commercial software package, in either one or two dimensions.
The results, for an assumed number of N = 100 contributing sta-
tionary phase paths, are shown in figures 5 and 6 for one- and two-
dimensional Gaussian images, respectively.
Figures 5 and 6 show a much richer structure than is evident in
their ensemble-average counterparts, figures 1 and 3, respectively,
although the overall power distribution in the snapshots appears
broadly consistent with that in the ensemble average cases - as it
should, of course. Of particular interest is that both snapshots man-
ifest inverted parabolic features; we shall refer to these features as
"arclets". The arclets can be understood in the following way. For
the linear image, the N speckles are characterised by their locations
on the x-axis: θi, i = 1...N, with the interference appearing in the
secondary spectrum at qij = θi − θj, pij = θ2
j . If we consider
a fixed value of i and allow j to vary, then we recognise that the
largest value of p corresponds to j = k such that θk < θj for
all j 6= k, so that θk lies close to the origin. For this point we have
qik ≃ θi, pik ≃ θ2
i , lying close to the apex of the inverted parabolic
arclet. Relative to this point, we can determine the locations of the
other interference terms, by noting that
i − θ2
qij − qik = −θj,
pij − pik = −θ2
j ,
whence
pij − pik = −(qij − qik)2,
(49)
(50)
which describes the locus of points in the ith arclet as the index j
is allowed to vary. Alternatively, if we vary i and keep j fixed then
the locus of points corresponds to
pij + θ2
j = (qij + θj )2,
(51)
which is an "upright" parabola of the same curvature (up to a sign)
as that of the individual arclets. Hence where there are gaps in the
image – i.e. no stationary phase points around that value of θ – there
are corresponding gaps in the intensity of the arclets, and those
gaps form parabolae cutting through the secondary spectrum. The
particular case where j = k in equation 51 describes a locus very
close to that of the apexes of the arclets: pik ≃ θ2
ik, so this
locus also exhibits the same curvature, but passes through the origin
of the secondary spectrum (i = k). Indeed the self power for all
image points (j = i), appears at q = 0, p = 0, so all of the
arclets pass through the origin. These features are strikingly similar
to what is seen in some of the data (see §6).
ysis leading to equation 50 yields instead the locus
pij − pik = −(qij − qik)2 − θ2
yj.
For the case of a snapshot of an axisymmetric image, the anal-
i ≃ q2
(52)
Figure 5. Secondary spectrum (delay-Doppler diagram) for a linear Gaus-
sian image in the speckle regime (with ψ = 0, so the image is elongated
along the direction of v⊥). As for the earlier figures, the axes here are given
in units of θo and θ2
o, for q and p, respectively. The expected image profile
is given by equation 39, and the actual image profile is shown in the inset
at the top of this figure. The ensemble-average secondary spectrum for this
circumstance is shown in figure 1.
Figure 6. As figure 5, but for the case of an axisymmetric Gaussian im-
age (as per equation 47) in the speckle regime; the actual speckle image is
shown in the inset box. The ensemble-average secondary spectrum for this
circumstance is shown in figure 3.
screens – and are thus appropriate to the long-term time-averaged
value of observations of secondary spectra. As discussed in §3, in-
dividual observations of a real source are not expected to be in this
regime. Usually the results which are reported lie in the average-
image regime – for which the model developed in §3 is still appro-
priate, albeit with the caveats given there – but in some instances
the data may be in the snapshot regime. In this case there is only a
single realisation of the randomness in the phase screen contribut-
ing to the observed secondary spectrum, and the phase screen can
be represented by a particular set of stationary phase points whose
specific locations are unknown. Obviously the snapshot secondary
spectra cannot be predicted in detail, but it is important to under-
stand their general appearance. We give two examples of snapshot
secondary spectra, for the cases of linear and axisymmetric Gaus-
sian images; these results may be compared directly with those of
§3.2.1 and §3.3.1, respectively.
8 Walker et al.
yj ≪ θ2
For speckles lying close to the x-axis we have θ2
xj, and these
points all lie close to the inverted parabolic arclet which defines
the upper envelope of power for interference with the ith point:
pij−pik = −(qij−qik)2. Referring to figure 6 we see that there are
indeed many points lying below each arclet, but the arclets them-
selves remain clearly visible in many cases. This is because the
relative delay of the interfering speckles is insensitive to the value
of θyj in the region θ2
xj, thus leading to a high density of
power close to the parabolic upper envelope of the arclet. This is
the same effect, described in §3.3.1, which leads to the visibility of
the single parabolic arc in figure 4. Similarly, the apexes of the ar-
clets do not follow a parabolic locus in the case of an axisymmetric
image; instead they are described by pik = q2
yi, so p = q2
ik + θ2
forms a lower envelope on the possible apex locations.
yj ≪ θ2
4.2 Evolution of features in the snapshot regime
The arclets shown in figures 5 and 6 are loci of the interference
terms between a given stationary phase point (speckle in the im-
age) and the line θy = 0. (More generally, for a linear image with
ψ 6= 0, the arclets in figure 5 would correspond to interference with
points close to the locus θy = θx tan ψ.) In the "frozen screen" ap-
proximation, the structure of the phase screen does not evolve, and
consequently individual stationary phase points simply drift across
the image as the line-of-sight moves relative to the screen (Melrose
et al 2004). Correspondingly, the apex of each of the arclets seen
in figures 5 and 6 is expected to drift across the (q, p) plane. In the
case of an isotropic image the apex of the ith arclet drifts along the
parabola p = q2 + θ2
yi with q ≃ v⊥/Ds and θyi = const.
In the case of a highly anisotropic image making an angle ψ = 0
with respect to the x-axis, the motion of the apex of the arclets
is along the parabola p = q2 with q ≃ v⊥/Ds. However, if
ψ 6= 0 then the misalignment between the linear image and the ve-
locity vector means that the stationary phase points cannot remain
fixed (even approximately) within the screen. Instead, each con-
tributing stationary phase point must slide in a direction perpendic-
ular to the image so that at every instant these points constitute a
linear image. Thus for highly anisotropic scattering we expect the
apex of each arclet to move along the parabola p = q2 sec2 ψ, with
q ≃ cos2 ψ v⊥/Ds.
Motion of the stationary phase point within the screen, which
occurs in the case of highly anisotropic scattering with ψ 6= 0, is
of interest because it implies a short lifetime for the corresponding
features in the secondary spectrum (even in the "frozen screen" ap-
proximation). The speed of motion of a given stationary phase point
relative to the phase structure in the screen is v⊥ sin ψ, in a direc-
tion perpendicular to the image elongation. If the coherence length
measured in this direction is So – which, by definition, is much
greater than so, the coherence length measured along the direction
of image elongation – then the corresponding feature in the sec-
ondary spectrum is expected to have a lifetime of ∼ So/v⊥ sin ψ,
and this may be very short unless ψ is very close to 0. We note that
in the development in §3.2, the limit So → ∞ was employed for
simplicity. This is, of course, an idealisation and in practice any
So ≫ so suffices to create highly anisotropic scattering.
lifetime of stationary phase points in highly
anisotropic scattering with ψ 6= 0 suggests that there may be an
observational bias favouring scattering screens with ψ ≃ 0, be-
cause these are the screens which yield persistent high-definition
arclets. (There may also be further observational biases favouring
ψ ≃ 0 in the case of scattering screens of finite spatial extent, be-
The short
Figure 7. Secondary spectrum (delay-Doppler diagram) for a multiple-
component, linear Gaussian image, as described by equation 53. As in fig-
ures 1–6, q is shown here in units of θo, and p in units of θ2
o.
cause of the longer interval over which such screens can contribute
to the received radiation.)
The foregoing considerations relate to the evolution of fea-
tures in the secondary spectrum under the assumption of a phase
screen which is "frozen", i.e. whose phase profile has no explicit
temporal evolution. However, the frozen approximation may be
poor over the time-scale taken for a given point to move right across
the image. If the phase profile of the screen evolves as a result of in-
ternal motions or wave-modes with velocity dispersion σ, then the
structure of a coherent patch, of size so, is expected to evolve on a
timescale ∼ so/σ and this gives us an upper limit on the lifetime
of any feature arising from a single stationary phase point.
5 DISTINCT IMAGE COMPONENTS
To model cases where there are multiple image components, dis-
tinct from the main image centred on θ = (0, 0), the anisotropic
image model of §3.2 is suitable. Here we consider the simplest mul-
tiple component configuration, in which there is only one additional
image, and this component has the same scattering properties (same
characteristic width, θo) as the main image; we have chosen an im-
age with ψ = 0 and
h1(θ) =
99
100
h(θ) +
1
100
h(θ − 10θo),
(53)
where h is the Gaussian profile given in eq. 39. In this case the ad-
ditional image component is weak (1% of the main component) and
centred on θx = 10θo; figure 7 shows the secondary spectrum for
this configuration. Compared with figure 1, the secondary spectrum
for a single-component linear Gaussian image, there are two new
features in figure 7: the thin linear feature emerging from the origin,
and the elongated blob of power with a peak at q = 10, p = 100.
The presence of the latter peak is expected, based on the analysis
presented for speckles in §4, as it corresponds to the central portion
of the weak component interfering with the central portion of the
main component. However, in contrast to the thin arclets present in
the snapshot example of figure 6, we see that the image represented
by eq. 53 yields a broad swath of extra power extending both to
larger and smaller p, q.
We can understand figure 7 in terms of the snapshot proper-
ties (figure 6) by considering the additional image component to
be made up of a large number of speckles, each of which yields a
thin arclet but with the apex at a different location for each speckle;
it then becomes clear that it is the angular extent of the additional
image which causes the power to be smeared out in the secondary
spectrum. Recall that the angular extent was chosen to be the same
as that of the main image component (eqs. 39, 53), so the main im-
age and the additional component should exhibit the same spread
in Doppler shift (q), but the quadratic mapping between image co-
ordinates and delay (p) means that the additional image component
exhibits a very extended delay spread because it is not centred on
the origin. It is helpful to illustrate this numerically, considering
only the central region of each image component: for the main im-
age component, extending over θx = (0 ± 1)θo, the delay (relative
to (0, 0)) ranges over 0 to 1, whereas for the additional image com-
ponent θx = (10 ± 1)θo with a delay range of 80 to 120. This
large spread is directly responsible for the large extent in p of the
additional power in figure 7.
The fact that the delay varies linearly across the off-centre im-
age component means that its self-interference term differs from
that of the main component, even though the two components have
the same shape. At the centre of the additional image component,
we have ∂p/∂q = 2q = 20, so that the self-interference of this im-
age component resides in a linear feature sprouting from the origin
of the secondary spectrum, close to the line p = 20q .
Although real data do sometimes show an extended swath of
power in the secondary spectrum, spread over a large range of de-
lays, it is more common to see a number of thin arclets. If these
arclets are transient (see §4.2) then they may be explicable in terms
of image speckle (§4), but if not then we need to accommodate
them within the context of distinct multiple image components. We
have seen (§4) that highly anisotropic images yield thin arclets, be-
cause there is a unique relationship between p and q when we con-
sider interference with a given speckle. However, when we consider
the addition of an extended, off-centre image component, even in
the limiting case of a linear image, the arclet will exhibit a thick-
ness (diameter) ∆p ∼ 4q∆θ, where ∆θ is the angular diameter
of the additional image component. The thickness of the arclet will
be greater than this if the anisotropy of the core of the image is
less extreme. The existence of persistent, thin arclets with apexes
at q ≫ 1 would therefore place strong constraints on the model
we have utilised, requiring that the additional image components
introduced to explain the arclets are subject to much less angular
broadening than the core of the image.
6 DISCUSSION
How well do our statistical models (§3) match the observations?
For the dozen bright pulsars studied by Stinebring et al (2004),
scintillation arcs or related phenomena are detected in all of them.
This effectively rules out an axisymmetric Gaussian intensity pro-
file (Figure 3) since it does not have sufficient power in the wings
of the profile to produce scintillation arcs. (The power distribution
near the origin of the secondary spectrum can be produced by ei-
ther a Gaussian or a Kolmogorov profile since they are similar for
θ 6 θ0.) A Kolmogorov profile with some degree of anisotropy in
the image is a much better fit to most of the secondary spectra. For
example, in Figure 8 we show a secondary spectrum for a pulsar
(PSR B1929+10) that consistently shows well-defined scintillation
arcs with a smooth, symmetric fall-off of power from the origin.
This appears to match best with some form of Kolmogorov image
(Figure 2 or 4). Determining the degree of asymmetry in the im-
age will require a more extensive comparison of model and data
Parabolic arcs
9
Figure 8. The secondary spectrum of PSR B1929+10 observed at 321 MHz
at Arecibo at epoch 2003.65. Contour levels are spaced at 3 dB intervals, as
for the theoretical results shown in figures 1–4, 7; the lowest level contour
is –45 dB below the maximum. The observed frequencies (ft, fν ) have
been converted to (q, p) (as per eqns. 28, 29), and scaled to correspond to
the axes in figs. 1–7, as described in the text. To do so we made use of
the measured values Dp = 0.33kpc vp = 163km s−1 for this pulsar,
together with the assumptions that vp dominates the contribution to v⊥
and β = 0.37, and θo ≃ 0.50 mas, which are estimated from the main arc
curvature (Stinebring et al 2001) and measured decorrelation bandwidth,
respectively.
than we attempt here. A visual comparison of the contour plot with
Figure 4, however, suggests that this image may be close to being
axisymmetric.
The most striking secondary spectra are those with pro-
nounced substructure, particularly those exhibiting inverted arclets.
As we discussed in §4, these can be reproduced by a model in which
point-like image maxima interfere with broader features near the
origin of the image. In Figure 5, for example, the separation be-
tween isolated image maxima causes gaps, particularly along the
delay (p) axis, between the inverted arclets. As shown in Figure 9,
there is remarkable qualitative agreement between observations and
the simple point-interference model of §4. The arclets in Figure 9
are clearly delineated (thin), have apexes that lie along a p ∝ q2 arc
that has the same (absolute value of) curvature as the arclets, and
have one side that passes through the origin. All of these features
are consistent with our model. Notice that the power asymmetry of
the arclet (extending further to the right than to the left) is the mir-
ror image of the power distribution along the main arc, where the
arc has greater power for −q values, but extends out further on the
+q side of the plot. This is to be expected if, as in our model, the ar-
clet is the interference between a point-like feature and the central
portion of the image on the sky. The observation shown here has
numerous counterparts for this pulsar over several years of recent
observations, and there is at least one other pulsar (PSR B1133+16)
that exhibits pronounced arclets at numerous epochs.
The models in §4 were constructed to describe the effects of
image speckle in the snapshot regime, and the evolution of the cor-
responding features in the secondary spectrum was discussed in
§4.2. The timescale on which the arclets in B0834+06 change is
observed to be very long, in some cases at least. In recent observa-
tions (Hill et al 2004), a persistent arclet pattern has been seen in
B0834+06 for more than 25 days. (This is much longer than either
the diffractive timescale, of 2 minutes, or the refractive timescale,
of 2.6 days, for this pulsar at the observing frequency of 321 MHz.)
10 Walker et al.
Figure 9. The secondary spectrum or PSR B0834+06 observed at 321 MHz
on 2004 Jan 8 with the Arecibo telescope. The greyscale is logarithmic in
relative power, with white being set 3 dB above the noise floor and black be-
ing set at 5 dB below maximum power. The arclet pattern seen prominently
on the right hand side of the plot persisted for more than 25 days during
2004 Jan and moved systematically up and to the right along the guiding
parabolic arc shown by the dashed line. The axis scaling is described in the
text and can be compared directly with the plots of model data. The axes
in this plot are scaled in the same way as in figure 8, but using the values
Dp = 0.72kpc, vp = 175km s−1, β = 0.33, and θo ≃ 0.72 mas.
Furthermore, the pattern shifts systematically in the direction from
−q to +q in a linear fashion ( q = constant) at a rate consistent
with the proper motion of the pulsar (v⊥ ≃ (β − 1)vp). These
properties point to a phase profile that is localised in one region of
the scattering screen and is scanned by the pulsar as it moves across
the sky; this circumstance is consistent with the evolution expected
in the case of a "frozen screen", as discussed in §4.2. However, the
large delays associated with the observed arclets (up to 300 µs) re-
quire the phase coherence length to be as small as so ∼ 109 cm,
if they are caused by diffraction. If we assume that the character-
istic velocity dispersion in the phase screen is σ & cs, the sound
speed, then with cs ∼ 10 km s−1 we expect that individual sta-
tionary phase points should evolve on a time-scale . 103 s. This
is 1/2000 of (the lower limit on) their observed lifetime, and we
conclude that the observed features cannot be attributed to isolated,
individual image speckles,1despite the superficial similarity of the
data to figure 5. Instead these features must be attributed to either (i)
a collection of many stationary phase points, if the corresponding
image feature has a diffractive origin, or (ii) one or more stationary
phase points associated with strong refraction. In the former case
one can imagine that the physical cause might be a localised re-
gion of strong turbulence, in which the individual stationary phase
points come and go; the lifetime of the secondary spectrum fea-
ture is then determined by the decay of the turbulence and may be
much longer than that of any individual stationary phase point. On
the other hand, if the feature is associated with strong refraction
– i.e. a lens – then the relevant length scale in the phase screen
is ≫ rF ≫ so and the corresponding timescale for evolution of
the phase screen is naturally much longer. Indeed, in the case of
a lens the phase structure need not be stochastic and might be in
a quasi-steady state with no detectable evolution (e.g. Walker and
Wardle 1998). In either case the image component must be tightly
1 Strictly we should be considering isolated pairs of stationary phase points
(Melrose et al 2004), but the essence of the argument remains the same.
concentrated in order to avoid smearing the power over a range of
delays (cf. figure 7). These compact refracting/scattering structures
may be the same structures which are responsible for the Extreme
Scattering Events (Fiedler et al 1987, 1994).
Despite the lack of agreement with the snapshot model, the re-
sults of §4 are useful in understanding the inverted arclets. A com-
parison of Figure 9 with Figures 5 and 6 indicates that the image
must be highly elongated in order to produce the observed arclet
pattern. This is a general result. No inverted arclet observations
exist that are consistent with an axisymmetric image. Instead, the
presence of inverted arclets is always accompanied by an absence
of power in the q = 0, p > 0 region, and the vertices of the arclets
lie along or very near a definite p = ηq2 locus, with η constant
over many years. All of these aspects follow naturally from an im-
age that is highly elongated in one dimension and has numerous
point-like features that self-interfere to give the observed arclets.
Furthermore, the constancy of η indicates that the region of the
interstellar medium being probed has a consistent directionality as-
sociated with it, as would be produced by anisotropic scattering in
the presence of a magnetic field (Higdon 1984, 1986; Goldreich &
Sridhar 1995). This topic is explored further in Hill et al (2004).
One other insight emerges from a comparison between the
models presented here and observations of scintillation arcs (e.g.
Stinebring et al 2004). Interference between two identically scat-
tered image components (e.g. Figure 7) is rare. Our modeling
shows that the interference between two identically-scattered com-
ponents would spread into a broad swath along the scintillation
arc due to the wide range of differential delay values present. This
is rarely seen in existing data, with thin arclets or pointlike inter-
ference features being much more common. This important item
needs further exploration.
It is evident from the development in §§2–4 that secondary
spectra encode information on the structure of radio images (i.e.
radio-wave propagation paths) of pulsars. A great variety of struc-
ture is seen in observed secondary spectra, the vast majority of
which remains to be decoded. Figures 5, 6 and 9 demonstrate that
secondary spectra with thin arclets or other pointlike interference
are particularly rich in information. Referring to equation 15, we
see that N points in the image should yield N (N −1)/2 features in
the secondary spectrum. Consequently we have O(N 2) constraints
but only O(N ) unknowns, so if N ≫ 1 there is enough infor-
mation to deduce the image structure from the observed secondary
spectrum. Each stationary phase point (or point-like structure in
the scattering screen) yields a measurement of the phase (up to an
overall additive constant), phase gradient, and the combination of
second derivatives which determines the magnification (eqs. 5–7);
all of these quantities being determined at the location xi of the im-
age point. To the extent that the N image points sample the image
plane, these measurements therefore allow us to deduce the struc-
ture of the phase screen, φ. We emphasise that this analysis does
not require the assumption that the delays are purely geometric; the
large number of available constraints should in fact provide power-
ful tests of any approximations which are used in the inversion. The
main obstacle is observational: we require secondary spectra that
show strong arclets or other sharply delineated interference features
with good signal-to-noise ratio throughout. The data shown in fig-
ure 9 appear suitable, but we have not yet attempted the inversion.
7 CONCLUSIONS
Approximating the Fresnel-Kirchoff Integral by a sum over the sta-
tionary phase points of an arbitrary phase screen has allowed us to
develop a probabalistic description of the power distribution in pul-
sar secondary spectra. We have presented a single-integral formula-
tion of the power distribution for general axisymmetric images, and
a closed form expression for highly anisotropic (i.e. linear) images.
These results are descriptions of the (ensemble-)average properties,
but our development makes it obvious how to generate examples of
snapshot secondary spectra using the Monte Carlo method, and we
have illustrated this technique with two examples. Our results clar-
ify the origins of the parabolic arcs in observed secondary spectra,
and provide an explanation for the inverted parabolic arclets which
are also sometimes seen. The observed arclets are due to struc-
tured, highly elongated images. Long-lived arclets require corre-
spondingly durable features in the phase-screen, and the data show
that compact, isolated clumps of refracting/scattering material must
exist in the interstellar medium. Consideration of the lifetime of
highly anisotropic image components highlights an observational
bias in favour of those which are elongated along the direction of
the effective velocity vector.
8 ACKNOWLEDGMENTS
We have benefitted greatly from helpful discussions with Barney
Rickett and Jim Cordes.
REFERENCES
Armstrong J.W., Rickett B.J., Spangler S.R. 1995 ApJ 443, 209
Cordes J.M., Rickett B.J., Stinebring D.R., Coles W.A. 2004 (In
preparation)
Cordes J.M., Weisberg J.M., Boriakoff V. 1985 ApJ 288, 221
Ewing M.S., Batchelor R.A., Friefeld R.D., Price R.M.,
Staelin D.H. 1970 ApJL 162, L169
Fielder R.L., Dennison B., Johnston K.J., Hewish A. 1987 Nature
326, 675
Fielder R.L., Dennison B., Johnston K.J., Waltman E.B., Si-
mon R.S. 1994 ApJ 430, 581
Goldreich P., Sridhar S. 1995 ApJ 438, 763
Goodman J., Narayan R. 1989 MNRAS 238, 995
Gradshteyn I.S., Ryzhik I.M. 1965 Table of integrals, series and
products
Gupta Y., Rickett B.J., Lyne A.G. 1994 MNRAS 269, 1035
Gwinn C.R., Britton M.C., Reynolds
J.E., Jauncey D.L.,
King E.A., McCulloch P.M., Lovell J.E.J., Preston R.A. 1998 ApJ
505, 928
Harmon J.K., Coles W.A. 1983 270, 748
Hewish A. 1980 MNRAS 192, 799
Hewish A., Wolzszczan A., Graham D.A. 1985 MNRAS 213, 167
Higdon J.C. 1984 ApJ 285, 109
Higdon J.C. 1986 ApJ 309, 342
Hill A.S., Stinebring D.R., Barnor H.A., Berwick D.E., Webber A.
2003 ApJ 599, 457
Hill A.S. et al 2004 (In preparation)
Lee L.C., Jokipii J.R. 1975 ApJ 201, 532
Lee L.C., Jokipii J.R. 1976 ApJ 206, 735
Melrose D.B, Macquart J.-P., Zhang C.M., Walker M.A. 2004
MNRAS (Submitted)
Narayan R., Hubbard W.B. 1988 ApJ 325, 503
Parabolic arcs
11
Rickett B.J. 1990 ARAA 28, 561
Rickett, B.J., Lyne A.G., Gupta Y. 1997 MNRAS 287, 739
Roberts J.A., Ables J.G. 1982 MNRAS 201, 1119
Shishov V.I. 1974 Sov Astron AJ 22, 544
Stinebring D.R., McLaughlin M.A., Cordes J.M., Becker K.M.,
Espinoza Goodman J.E., Kramer M.A., Sheckard J.L., Smith C.T.
2001 ApJL 549, L97
Stinebring D.R. et al 2004 (In preparation)
Walker M.A., Wardle M.J. 1998 ApJL 498, L125
Wolszczan A., Cordes J.M. 1987 ApJL 320, L35
APPENDIX A: EVALUATION OF INTEGRALS FOR
AXISYMMETRIC IMAGES
In evaluating the integral
1 + θ2
P =Z d2θ1d2θ2 p(θ1)p(θ2) δ(p − θ2
2)δ(q − θ1x + θ2x),
(A1)
one may assume p > 0 without loss of generality, with the func-
tion for p < 0 determined from that for p > 0 by P (−p, q) =
P (p,−q). For the axisymetric images under consideration here, the
secondary spectrum is also symmetric in q, and in this Appendix we
further limit the domain of our calculation to q > 0. One may carry
out any two of the four integrals in eq. A.1 over the δ-functions.
We evaluate the integrals in polar coordinates, d2θi = θidθidψi.
Carrying out the integral over d2θ1 gives
P (p, q) =X± Z d2θ2
p(θ1)p(θ2)
2θ1y
(cid:12)(cid:12)(cid:12)(cid:12)θ1y =±Y, θ1x =q+θ2x
,
(A2)
where θ1y is determined by the δ-functions, and is given by
Y 2 = p + θ2
2 − (q + θ2 cos ψ2)2.
(A3)
The boundary of the physical region is defined by Y = 0. This
corresponds to a parabola with apex at θ2x = (p − q2)/2q, and
asymptotes at θ2y → ±∞, θ2x → +∞. The integral is over the
region outside this parabola. For (p − q2)/2q > 0 the apex of
the parabola is to the right of the origin, and the integral is over
0 6 ψ2 < 2π for θ2 < (p − q2)/2q, and over cos ψ2 6 cos ψ+,
where
cos ψ+ = pp + θ2
θ2
2 − q
,
(A4)
for θ2 > (p−q2)/2q. For (p−q2)/2q < 0 the apex of the parabola
is to the left of the origin, so that a region around the origin is not in
the range of integration, and the integral is over cos ψ2 6 cos ψ+
for θ2 > (p − q2)/2q.
change variables from ψ2 to t, using
One procedure for evaluating the angular integral is to first
t = tan
ψ2
2
,
Then one finds
dψ2 =
2dt
1 + t2 ,
cos ψ2 =
1 − t2
1 + t2 .
(A5)
Y =
(p − q2 + 2qθ2)1/2
(1 + t2)
where
(cid:2)(cid:0)t2 + a2(cid:1)(cid:0)t2 + b2(cid:1)(cid:3)1/2,
b2 = pp + θ2
2 − q − θ2
pp + θ2
2 − q + θ2
.
a2 = pp + θ2
pp + θ2
For p > q2 and 0 < θ2 < (p − q2)/2q (with q > 0), we have
2 + q + θ2
2 + q − θ2
(A7)
,
(A6)
12 Walker et al.
a2 > b2 > 0, while for θ2 > (p − q2)/2q we have b2 < 0 < a2.
Introducing d2 = −b2, and making use of Gradshteyn and Ryzhik
(1965: 3.152, 8.111.2, 8.112.1, 8.113.1) we find that the requisite
integrals are
and
0
Z ∞
Z ∞
d
dx
[(a2 + x2)(x2 + b2)]1/2 =
1
a
(cid:19) ,
(A8)
a
K(cid:18)√a2 − b2
K(cid:18)
1
dx
[(a2 + x2)(x2 − d2)]1/2 =
√a2 − b2
a
√a2 − b2(cid:19) ,
(A9)
where K is the Complete Elliptic Integral, with
K(k) =
π
2 (cid:18)1 +
k
4
+ . . .(cid:19) .
The foregoing results imply
for 0 < θ2 < (p − q2)/2q (with q > 0), and
Z π
0
dψ2
Y
=
1
√A
K(cid:16)pB/A(cid:17) ,
Z π
ψ2+
dψ2
Y
=
1
√
B
K(cid:16)pA/B (cid:17) ,
for θ2 > (p − q2)/2q, where
A ≡(cid:18)qp + θ2
2 + θ2(cid:19)2
− q2
and
(A10)
(A11)
(A12)
(A13)
(A14)
B ≡ 4θ2qp + θ2
2.
Substituting these results into equation A2 then yields our final re-
sult for the secondary spectrum corresponding to an axisymmetric
image (equation 43).
|
astro-ph/0506321 | 1 | 0506 | 2005-06-14T20:04:13 | Radiative Transfer and Acceleration in Magnetocentrifugal Winds | [
"astro-ph"
] | Detailed photoionization and radiative acceleration of self-similar magnetocentrifugal accretion disk winds are explored. First, a general-purpose hybrid magnetocentrifugal and radiatively-driven wind model is defined. Solutions are then examined to determine how radiative acceleration modifies magnetocentrifugal winds and how those winds can influence radiative driving in Active Galactic Nuclei (AGNs). For the models studied here, both radiative acceleration by bound-free (``continuum-driving'') and bound-bound (``line-driving'') processes are found to be important, although magnetic driving dominates the mass outflow rate for the Eddington ratios studied (L/L_Edd = 0.001 - 0.1). The solutions show that shielding by a magnetocentrifugal wind can increase the efficiency of a radiatively-driven wind, and also that, within a magnetocentrifugal wind, radiative acceleration is sensitive to both the column in the shield, the column of the wind and the initial density at the base of the wind. | astro-ph | astro-ph | ACCEPTED TO APJ
Preprint typeset using LATEX style emulateapj v. 11/26/04
5
0
0
2
n
u
J
4
1
1
v
1
2
3
6
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
RADIATIVE TRANSFER AND ACCELERATION IN MAGNETOCENTRIFUGAL WINDS
Canadian Institute of Theoretical Astrophysics, University of Toronto, 60 Saint George Street, Toronto, ON M5S 3H8, Canada
JOHN E. EVERETT
Accepted to ApJ
ABSTRACT
Detailed photoionization and radiative acceleration of self-similar magnetocentrifugal accretion disk winds
are explored. First, a general-purpose hybrid magnetocentrifugal and radiatively-driven wind model is defined.
Solutions are then examined to determine how radiative acceleration modifies magnetocentrifugal winds and
how those winds can influence radiative driving in Active Galactic Nuclei (AGNs). For the models studied here,
both radiative acceleration by bound-free ("continuum-driving") and bound-bound ("line-driving") processes
are found to be important, although magnetic driving dominates the mass outflow rate for the Eddington ratios
studied (L/LEdd = 0.001- 0.1). The solutions show that shielding by a magnetocentrifugal wind can increase the
efficiency of a radiatively-driven wind, and also that, within a magnetocentrifugal wind, radiative acceleration
is sensitive to both the column in the shield, the column of the wind and the initial density at the base of the
wind.
Subject headings: galaxies: active -- galaxies: jets -- hydrodynamics -- MHD -- radiative transfer -- quasars:
general
1. INTRODUCTION
A variety of observational signatures point to the im-
portance of outflowing gas within many types of Active
Galactic Nuclei (AGNs). Blueshifted absorption features
(in Broad Absorption Line Quasars, or BALQSOs; see,
e.g., Weymann et al. 1991) are seen in approximately 15%
(Reichard et al. 2003) of radio-quiet quasars, with velocities
up to 0.1c. In addition, radio-loud quasars display relativis-
tic, collimated outflows. More recently, both UV and X-ray
absorbing gas have been observed in approximately 60% of
Seyfert 1 galaxies (Crenshaw et al. 1999). Observational es-
timates hint that the mass outflow rate is nearly equal to the
mass inflow rate (for a review of mass outflow in AGNs, see
Crenshaw, Kraemer, & George 2003).
The development of models to explain the mass outflow
rates, geometry, and general kinematics of these winds has
proven difficult, but progress on a number of possibilities has
been encouraging. From very early on, researchers exam-
ined both radiative wind models (e.g., Drew & Boksenberg
1984; Vitello & Shlosman 1988) and, to explain radio jets,
hydromagnetic models (e.g., Blandford & Payne 1982, here-
after BP82). Later models of radiatively-driven winds were
able to explain the BALQSO outflow velocities and popu-
lation fraction (Murray et al. 1995, hereafter, MCGV95) as
well as the single-peaked emission lines (Murray & Chiang
1997).
The line-driven models were also demonstrated
in hydrodynamical simulations (Proga, Stone & Drew 1998;
Proga, Stone, & Kallman 2000; Pereyra et al. 2004), which
elucidated the density structure, geometry, and kinematics
of the flow in two dimensions.
In addition, models with
combinations of continuum- and line-driving by X-rays were
presented by Chelouche & Netzer (2001). One of the con-
cerns with line-driving in AGNs has been the possible over-
ionization of the gas by X-rays in the central continuum
(leaving the wind with too few lines to intercept flux in
the lines): the need for "shielding gas" to prevent this ove-
rionization has been a persistent concern (e.g., MCGV95,
Chelouche & Netzer 2003b; Proga & Kallman 2004).
Electronic address: [email protected]
In addition, hydromagnetic winds have also been devel-
oped to gain insight into the observations; some of these
models have included radiative driving.
In these magneto-
centrifugal winds, gas is loaded onto and tied to large-scale
magnetic field lines; those field lines then fling the matter
centrifugally away from the disk, like beads along a wire.
Magnetocentrifugal acceleration is commonly used to ex-
plain large-scale collimated outflows in radio galaxies (BP82;
for reviews of magnetocentrifugal driving, see Spruit 1996;
Königl & Pudritz 2000; Ferreira 2003). In the context of the
Broad Emission Line Region (BLR), magnetocentrifugal out-
flows made up of clouds were first called upon to explain
the single-peaked broad emission lines, outflow velocities,
and densities (Emmering et al. 1992) . The "torus" (obscur-
ing gas that plays a central role in the Unification model,
yielding a dependence of observed properties with inclination
angle; see Antonucci 1993) has been explained as a dusty,
continuous magnetic wind (Königl & Kartje 1994, hereafter
KK94) where radiative acceleration on dust affects the wind
geometry. Hydromagnetic disk winds with radiation pressure
were also developed by de Kool & Begelman (1995) as an al-
ternative explanation for the population fraction of BALQ-
SOs and to understand cloud confinement within the out-
flow.
In addition, another magnetocentrifugal wind model
was developed to explain single peaked emission lines aris-
ing from a distribution of clouds (Bottorff et al. 1997), as
well as the dynamics of the warm absorber in NGC 5548
(Bottorff, Korista, & Shlosman 2000). Finally, some effects
of magnetic fields (not including magnetocentrifugal driving
as in BP82) have also been considered in two-dimensional hy-
drodynamic simulations (Proga 2000, 2003).
At the present time, both radiatively driven winds and mag-
netocentrifugally driven winds (with radiative acceleration
added in some models) offer compelling but competing pic-
tures of the key physics in the cores of AGNs. No clear obser-
vational evidence yet discriminates the dominant physics of
wind-launching in AGNs. For radiatively-driven winds, three
main lines of evidence hint that radiative driving should be
important in models of wind dynamics. First, Laor & Brandt
(2002) find a correlation between UV luminosity and the ob-
2
Everett
served outflow velocity, which agrees with a basic predic-
tion of line-driven wind models (Proga 1999).
In addition,
the "ghost of Ly-α" (Arav et al. 1995; Chelouche & Netzer
2003b), as well as the realization that the radiative momen-
tum removed from the continuum in blueshifted absorption
lines in BALQSOs is a significant fraction of the total ra-
diative momentum (MCGV95) are both also important clues
that radiative acceleration is an important component to any
self-consistent model of AGN winds. There are still, how-
ever, concerns about how the "shielding" of the wind works
(Chelouche & Netzer 2003b; Proga & Kallman 2004). In ad-
dition, in the case of stellar disk winds, observations of two
nova-like variables seem to show that their winds are not dom-
inated by radiative driving (Hartley et al. 2002). (This con-
clusion, however, rests on the prediction that line equivalent
widths are direct measures of mass outflow rate, which may
not be the case; see Pereyra et al. 2004)
On the other hand, the leading model for collimated radio
jets in AGNs (BP82) already calls upon large scale, dynami-
cally important magnetic fields, as do the hydromagnetic wind
models (mentioned above, e.g. Königl & Kartje 1994; Kartje
1995; Bottorff et al. 1997) that have also had success in ex-
plaining AGN observations. In addition, such a hydromag-
netic wind would have no difficulty with overionization, and
so might naturally serve as a "shield" for radiative accelera-
tion further from the central source. There are, however, cur-
rently no models which address the interplay between magne-
tocentrifugal driving and radiative acceleration; if such wind
models could be constrained, we may be able to observation-
ally distinguish the physics of wind launching in the cores of
AGNs, and gain insight into the role that outflows play both
in accretion and in feedback of those winds in the galaxy and
surrounding matter.
This paper develops a detailed photoionization and dynami-
cal model for magnetocentrifugal winds in AGNs. The model
is designed to explore the radiative transfer within such mag-
netocentrifugal winds, but also to help understand how radia-
tive driving impacts the kinematic structure of such winds.
Constructing such a model builds the foundation for later
work to determine absorption and emission line profiles in or-
der to compare with observations and check for the presence
of magnetocentrifugal winds within AGNs. An overview of
this model is presented in §2; the model is then defined in
detail in §3. An examination of the structure of a particular
"fiducial" magnetocentrifugal, radiatively-accelerated wind is
described in §4 and then the dependences of radiative acceler-
ation on some initial parameters are shown in §5. Conclusions
and directions for future work are summarized in §6.
2. MODEL OVERVIEW
Before examining the model in detail, it is instructive to
present a summary of the basic design. The semianalytic
model developed here includes magnetic acceleration and ra-
diative acceleration of a continuous wind launched from an
accretion disk. A detailed treatment of radiative transfer is in-
cluded by using Version 96.00 of the photoionization simula-
tion program Cloudy, last described by Ferland et al. (1998).
These elements are introduced in an approximate schematic of
the wind model shown in Figure 1, depicting a portion of the
accretion disk and outflowing wind. In this figure, radiation
(entering from the left side of the schematic) first encounters
a purely magnetocentrifugally accelerated wind, which will
be referred to as the "shield", as it absorbs radiation from
the central continuum. The shield is introduced as a sepa-
rate component in order to cleanly differentiate the effect of
shielding from radiative acceleration; radiative driving of the
shield is therefore not considered in this work. Beyond that
shield is an optically thin, radiatively and magnetically ac-
celerated wind streamline (which we will henceforth refer to
as the "wind"). In this portion of the model, both magneto-
centrifugal and radiative forces help accelerate the flow off of
the accretion disk; the magnetic fieldlines are shown by the
black lines bordering the outflow. The included radiative ac-
celeration is calculated by first simulating the photoionization
within both the shield and the wind along radial paths such as
the thick, black lines in the figure.
The separation of the wind into two components is, of
course, artificial. In reality, radiative acceleration would grad-
ually increase in importance for portions of the wind that
are increasingly shielded. However, splitting the outflow into
these two components allows a first-order, qualitative solution
that can be used to gain some understanding of how magnetic
and radiative forces might interact, and how a magnetic wind
may be able to act as a radiative "shield" to allow more ef-
ficient radiative acceleration. While artificial, this method of
using two wind components has already been used success-
fully to examine winds with magnetocentrifugal and radiative
driving on dust (e.g., KK94, Kartje 1995).
Figure 2 presents a schematic flow chart of how model cal-
culations proceed. The wind starts as a self-similar magneto-
hydrodynamic model that yields the pure magnetocentrifugal
wind solution (covered in §3.1). Next, simulations of the pho-
toionization balance of that wind streamline (§3.2) are run,
and the resultant ionization balance and transmitted contin-
uum are used to calculate the radiative acceleration behind
the shield (§3.3). Next, the radiative acceleration is input (as
a function of polar angle, θ) back into the self-similar magne-
tohydrodynamic model, modifying the structure of the wind
streamline, while leaving the shield unaffected. This process
is then repeated, simulating the photoionization of that mod-
ified wind and recalculating the radiative acceleration terms.
We typically iterate five to eight times to converge to a final
equilibrium solution.
With the basic model now summarized, it may be instruc-
tive to compare and contrast it to the recent wind model
examined in Proga (2003), where the combination of mag-
netic and radiative forces in disk winds is also investigated.
Proga (2003) concentrates on numerical simulations of time-
dependent winds with line driving and magnetic forces. These
numerical simulations allow large-scale models of outflows
that are valuable in understanding global wind structures
in many different astrophysical contexts.
In contrast, the
radiatively-driven components of this model are more local-
ized, since radiative acceleration is applied to the shielded
streamline only. This model setup has been chosen for its
flexibility in radiative acceleration modeling; using Cloudy
for such radiative acceleration models enables computations
not only of radiative line driving but also of continuum driv-
ing, and allows us the freedom to include dust and easily vary
the incident spectrum, for example. In addition, the magnetic
winds produced in Proga (2003) are not magnetocentrifugal
outflows (as in BP82), as are the semianalytic winds presented
here. Further, these new models are steady-state, not time-
dependent. Finally, semi-analytic, steady-state models allow
an exploration of general behaviors through many parameter
variations; large-scale numerical simulations can usually vary
only a few parameters. In summary, these models cover dif-
ferent facets of the disk-wind problem, yielding different per-
Radiative Acceleration of MHD Winds
3
B
"shield"
B
B
Radiatively Accelerated
Flowline
("wind")
Sample Cloudy
Simulation Zones
Accretion Disk
FIG. 1. -- Schematic of the basic geometry and major components of the radiatively accelerated magnetocentrifugal wind model. The wind itself is split
into two components: a pure magnetocentrifugal wind that acts as a shield (the wider component on the left) and an optically thin streamline with combined
magnetocentrifugal and radiative acceleration. The heavy black lines indicate a few of the radial zones where Cloudy simulates the photoionization balance of
the wind (typically, ∼ 40 such radial Cloudy simulations are run, spaced logarithmically in polar angle, θ).
Magnetic Field Lines
Magnetic Field Lines
Radiation
wind
shield
Accretion Disk
Accretion Disk
Radiation
Magnetic Field Lines
Accretion Disk
FIG. 2. -- The iterative scheme in the wind model, showing a single iteration of the model, moving counterclockwise around the figure. The model starts by
solving the self-similar magnetohydrodynamic (MHD) equations for the structure of a centrifugally driven outflow (upper left). Next, photoionization simulations
determine the continuum incident on the radiatively accelerated streamline as well as the ionization state of the gas in the streamline, and then those results are
utilized to calculate the radiative acceleration of that flow. The radiative acceleration is then applied (as a function of the polar angle θ) to the second component
of the self-similar magnetocentrifugal wind. A new wind structure for that second section of the outflow is then derived from the interaction of those forces
(upper right).
spectives on a complicated system.
3. THE TWO-PHASE HYDROMAGNETIC AND RADIATIVE WIND
MODEL
In this section, we describe in detail the model's com-
ponents, and derive key equations. The magnetocentrifugal
wind model is introduced first (§3.1), followed by the pho-
toionization simulations (§3.2), radiative acceleration calcu-
lation (§3.3), and finally a discussion of the wind model equa-
tion of motion (§3.4).
3.1. Magnetocentrifugal Self-Similar Wind Solution
To derive the equations governing the continuous magneto-
centrifugal wind, we start with the equations of a stationary,
axisymmetric magnetohydrodynamic (MHD) flow in cylin-
drical coordinates, make the assumption of self-similarity
in the spherical radial coordinate, and utilize the continuity
equation, conservation of angular momentum along the flow,
and both the radial and vertical momentum equations, very
much as in BP82 and KK94 (see Appendix A). Thermal ef-
fects are neglected in the wind, therefore effectively assuming
that the wind starts out supersonic (this assumption is checked
later on, see §3.5). In deriving the wind equations, we use the
same simplifications as BP82, except for the added complica-
4
Everett
tion that energy is not conserved in the radiatively-accelerated
system due to the constant input of radiative energy into the
wind. In the original formulation of BP82, conservation of
energy supplied an additional constraint which allowed a sim-
plification of the equations of motion to two first-order differ-
ential equations. Because energy is not conserved in this flow,
the equivalent of three first-order differential equations must
be integrated, solving for three parameters simultaneously in-
stead of two as in BP82. The detailed setup and derivation of
this set of equations of motion are given in Appendix A.
3.2. Photoionization Simulations of the Wind
Next, Version 96.00 of the photoionization code Cloudy
(Ferland et al. 1998) is used to simulate the absorption of the
magnetocentrifugal shield and wind as well as the ionization
state at the wind streamline. Photoionization simulations of
the shield and wind are used to calculate the radiative accel-
eration and allow considerable flexibility in gas parameters
(such as gas abundances, dust, central continuum, etc). This
flexibility is gained at the cost of simulating the photoioniza-
tion state of the wind as if it were a static medium, as Cloudy
assumes; this is only true of the recombination timescale for
the gas is much shorter than the transit timescale of gas in
the region simulated (τrecomb < τflow). We have however, veri-
fied, a posteriori, that τrecomb < τflow for all of the radiatively
accelerated ions at the base of the wind where radiative ac-
celeration is important, using the recombination rate approxi-
mations in Arnaud & Raymond (1992) and Verner & Ferland
(1996) to calculate the recombination times. Even at high lat-
itude, τrecomb < τflow for all of the significant ions. At the high-
est latitudes, highly ionized O, which has the longest recom-
bination time among the significantly radiatively accelerated
ions, has a recombination timescale that is still a factor of two
4πρvp
Bp
+ v2
The integration of the momentum equations starts by spec-
ifying the following initial parameters: the mass loading of
the wind (the ratio of mass flux to magnetic flux in the mag-
, where ρ is the mass den-
netocentrifugal wind, κ ∝
sity of the wind, vp is the poloidal velocity of the wind
(vp ≡ (v2
z )1/2), and Bp is the poloidal magnetic field
r
strength), the specific angular momentum of gas and field in
the wind, and the power-law exponent (b) that describes the
change in density with spherical radius: ρ ∝ R- b. Also in-
put, as parameters, are the mass of the central black hole, M•,
the wind's launch radius on the disk, r0, and the density at
the base of the wind at the launch radius, n0. The program
employs a "shooting" algorithm (using the SLATEC routine
DNSQ; Powell 1970) to integrate from the singular point (the
Alfvén point) to the disk, solving for the height of the singu-
lar point above the disk (χA) and the slope of the streamline at
both the disk and the Alfvén point (ξ ′
A) by matching the
integration results to boundary conditions on the disk. After
solving for the position of the Alfvén point, the equations of
motion are integrated from the disk to a user-specified height
beyond the Alfvén point; along the streamline, the run of ve-
locity, density, and magnetic field are calculated.
0 and ξ ′
This code has been tested (without radiative acceleration)
against the solution given in BP82 and have duplicated their
results to within 8%. This is close to the previously reported
4% variance in the recalculation of BP82 reported in Safier
(1993). The difference in these new results comes not only
from using higher precision calculations compared to BP82
(we use the same precision as Safier 1993), but in addition, a
more complex set of equations is being solved.
less than the transit time. Meanwhile, highly ionized Fe ions,
which dominate the low level of radiative acceleration at high
latitudes, have recombination timescales an order of magni-
tude less than the transit timescale. Using Cloudy is there-
fore a reasonable approximation for the radiative equilibrium
within our winds (especially since, with our code, adiabatic
and advective effects are added, see §3.2.1). As Cloudy en-
ables a flexible, self-consistent calculation of radiative accel-
eration, we accept this approximation to enable these calcula-
tions.
Of basic importance to the photoionization simulations
is the illuminating spectral energy distribution (SED). For
the purposes of these simulations, an SED adapted from
Risaliti & Elvis (2004) is used (an example of the SED is
shown in Figure 6). This SED is input into Cloudy via
the generic "AGN" continuum with Tblackbody = 1.5 × 105 K,
αox = - 1.43 (Elvis, Risaliti, & Zamorani 2002), αUV = - 0.44,
and αX = 0.9 (αox defines a single power-law that would de-
scribe the continuum between 2500 Å and 2 keV, αUV is the
slope of the low-energy component of the Big Blue Bump,
and αX is the X-ray power-law exponent; our value of αox
is taken from the middle of the range 0.8 to 1.0 given in
Risaliti & Elvis (2004)).
Spectral signatures and radiative acceleration also depend,
of course, on the column density in the wind. Observations
yield only rough constraints for this, so these columns as left
as free parameters; the effect of varying these columns will be
investigated in this paper. The columns throughout the shield
and wind are set by the columns at the base of the shield and
wind, denoted NH,0 for the hydrogen column at the base of
the wind. As the wind rises above the disk and accelerates,
that column density (NH) drops as a function of height due
to mass conservation (an example of this is shown later in
Fig. 7). Investigating the shielding ability of such a dynamic
shielding column is of central interest to this paper, and will
be addressed in §5.1.
3.2.1. Photoionization of the Continuous Wind
As depicted in Figure 1, Cloudy simulates the photoioniza-
tion of the wind along radial sight lines through the shield
and through the wind, ending at the site of radiative accel-
eration on the wind streamline. The continuum incident on
the wind streamline is also calculated. The photoionization
state and continuum at the end of the Cloudy calculations are
recorded, and then used to compute the radiative acceleration
of that gas. Finally, the acceleration is tabulated and applied
as a function of θ along the wind streamline by inputing the
angle-dependent radiative acceleration into the gravitational
term (this is covered in more detail in Appendix A, see eqn.
A3).
Whereas Cloudy is designed to simulate the photoioniza-
tion balance of gases as in the shield and wind, it cannot eas-
ily incorporate adiabatic and advection effects: Cloudy has
no knowledge of the particular velocity profile of the wind
in the overarching model, or the temperature difference be-
tween successive photoionization models as the wind climbs
above the disk. Therefore, this model calculates both advec-
tive heating and adiabatic cooling in the wind, and adds those
terms manually into Cloudy's simulations. Both terms largely
cancel in the wind, and have only a negligible effect on out-
flow dynamics, but they are included in all of the models for
completeness.
3.3. Radiative Acceleration Calculations
Radiative Acceleration of MHD Winds
5
The model then incorporates the above-mentioned results
for the ionization structure and radiation field to calculate the
radiative forces felt by the wind. There are two different kinds
of radiative acceleration to consider: continuum acceleration
(including radiative acceleration on dust) and line accelera-
tion. It is convenient to express the radiative acceleration in
terms of Γ(θ),
Γ(θ) ≡
aradiative(θ)
g
,
(1)
where aradiative is the acceleration due to radiation, and g is the
local gravitational acceleration.
3.3.1. Line and Continuum Acceleration
In general, for continuum and line acceleration, the radia-
tive acceleration is given by
Γ =
neσT F
ρc
(Mcont + Mlines)
GM•
r2+z2
,
(2)
where F is the local flux (the flux transmitted through both
the shield and wind column), ne is the electron density, ρ is
the gas density, c is the speed of light, G is the gravitational
constant, M• is the mass of the central black hole, r and z are
cylindrical coordinates centered on the black hole, and Mcont
& Mlines are the "force multipliers" that relate how much the
radiative forces on the gas (on the line and continuum opacity,
respectively) exceed the radiative forces on electrons alone.
They are given below in terms of the continuum opacity, χν,
and the line opacity, χl, for the continuum and lines, respec-
tively:
Mcont =
Mlines =
1
neσT F Z χνFνdν,
1 - e- ηlt
F Xl
Fl∆νl
1
,
t
with
ηl ≡
χl
σT ne
t ≡
σT nevth
dvR/dR
,
(3)
(4)
(5)
where ν is the photon frequency, Fl is the local (transmit-
ted) flux in the line at the frequency of line number l, vth
is the sound speed in the gas, ∆νl = νvth/c is the thermal
line width, and ηl compares the opacity of the line (for a
given ionization state of the gas) to the electron opacity, rep-
resenting all of the atomic physics in the radiative acceler-
ation calculation. The last remaining variable, t, is often
called the "effective electron optical depth" and encodes the
dynamical information of the wind in the radiative accel-
eration calculation. This dynamical information is impor-
tant because in an accelerating medium, one must also ac-
count for the Doppler shift of the atomic line absorption en-
ergy in the accelerating gas relative to the emitted line pho-
ton's energy: beyond the Sobolev length, vth/(dvR/dR), in-
cluded in t, a line photon will be Doppler-shifted out of
the thermal width of the absorption line and can escape
the gas (see Sobolev 1958; Castor, Abbott, & Klein 1975;
Mihalas & Weibel-Mihalas 1999).
The above-mentioned force multipliers are calculated us-
ing the resonance line data of Verner et al. (1996) with solar
abundances. Since Mcont, the continuum multiplier, depends
only on the ionization state, it is tabulated solely as a function
of height in the wind. In contrast to Mcont, the line multiplier
(Mline) is tabulated for a range of values of the parameter t.
Later, when calculating the equation of motion for the wind,
the local velocity gradient is used to compute the actual value
of t, which is then used to linearly interpolate the table of Mline
and then evaluate the radiative acceleration.
The force multiplier computation has been tested against
Arav et al. (1994), who also calculated radiative acceleration
from photoionization simulations. Figure 3 compares these
new calculations results against their fits (noting that there is
a typo in their eq. [2.9]; Z.-Y. Li, personal communication),
where the force multipliers as a function of the ionization pa-
rameter U is presented (U is the ratio of hydrogen-ionizing
photon density to hydrogen number density n, given by U ≡
Q/4πnR2c, where Q is the number of incident hydrogen-
ionizing photons per second, and R is the distance from the
continuum source). Overall, good agreement is found, espe-
cially considering that Arav et al. (1994) point out that their
fit deviates from their calculations at low values of U. The
increase in the newly-calculated continuum force multiplier
over Arav et al. (1994) is most likely due to the different con-
tinuum opacity database included in Cloudy 96 compared to
the code (MAPPINGS) that was used in Arav et al. (1994).
The multiplier values and trends with ionization parameter are
still clearly very similar, however.
3.3.2. Non-Sobolev Effects
Simply using the Sobolev length, vth/(dvR/dR), in Equation
4 can be misleading. In early simulations, we found that this
Sobolev length could, in regions where the gas is slowly ac-
celerating, be much larger than the physical size of the shield
and wind combined. This is clearly not physical, so a sim-
ple non-Sobolev method was employed to calculate the size
of the combined shield and wind column.
First, for the wind, the length of the absorbing column is
simply limited to the minimum of the wind's Sobolev length
and its true (physical) length.
Second, for the shield, the length of the absorbing column
is given by the minimum of the shield's Sobolev length1, its
true physical length, and the average length of the column
that absorbs photons at the wavelength of the dominant ac-
celerating atomic lines. This last length-scale requires more
explanation. To calculate this length, the top 20 line force
multipliers (for individual atomic lines) are found for each
polar angle θ, and for each of those lines, the column of each
particular ion in the shield is read from the Cloudy simula-
tions. An average shielding column for all of the high-opacity
lines is then calculated. To test this approximation, we have
used lists of the top 10, 20, and 50 transitions in the wind
and used them to calculate the limiting column in the shield.
Changing the number of lines included does not significantly
change the final wind solutions, especially since this effect is
only important at the extreme base of the wind. Without con-
sidering all of these constraints on the size of the absorbing
column, the Sobolev length can significantly overestimate the
optical depth in the shield (even overestimating the physical
length of the columns, for small accelerations), which results
in the line driving acceleration dropping below the continuum
acceleration. Other non-Sobolev effects (such as line blan-
keting within the shielding gas or the wind) are not consid-
1 The Sobolev length calculation for the shield does take into account the
offset in velocity between the wind and shield, which is important since the
wind is radiatively accelerated in addition to the magnetic acceleration that
the wind and shield share.
6
Everett
FIG. 3. -- Comparison between the force multiplier calculation in this code (solid lines) and the fits of Arav et al. (1994, dashed lines). The plot in the upper
left is the comparison of the continuum force multiplier, as a function of the ionization parameter, U. The next three panels show a comparison of the line force
multiplier (which is also a function of t, besides U) as a function of U for three different values of t, given at the top of each panel.
ered in this model; for a consideration of these effects, see
Chelouche & Netzer (2003a).
As we are interested in steady-state winds, ∂v
∂t in Eq. 6 is set
to zero. As already mentioned, gravitational, radiation, and
Lorentz forces are included, while thermal terms are not. This
yields the expression below, with the magnetic force term is
split into pressure and tension components.
ρ(v · ∇)v = - [1 - Γ(θ)]
GMρ
R2
R -
1
8π
∇B2 + 1
4π
(B · ∇)B (7)
For these calculations, it is more intuitive to integrate the
equation of motion along the flow already given by the mag-
netocentrifugal wind solution. Therefore, taking the dot prod-
uct of Euler's equation with s, which is defined as the direction
along the flow, and expanding and simplifying the left-hand
side of the equation, one finds:
(v · ∇v) · s = vp
∂vp
∂s
- v2
φ
r
sin θF ,
where
θF ≡ tan- 1(cid:18) dr
dz(cid:19) .
In the same way, if θ is defined via
θ ≡ tan- 1(cid:18) r
z(cid:19) ,
(8)
(9)
(10)
With a length-scale found from the sum of the wind and
shield lengths found above, that length-scale is then substi-
tuted for the Sobolev length in the calculation of Mlines in
Eq. 4. Including this physical length of the wind and shield
will introduce dependences on the sizes of the wind and shield
columns, which will be examined in §5.
3.4. Integrating the Euler Equation for the Wind
Given the results from the ionization and radiative accel-
eration calculations, the next step is to solve for the effect
of radiative acceleration on the wind, taking the magnetocen-
trifugal wind model already computed and augmenting its ac-
celeration with radiative forces. To do this, the full equation of
motion (the Euler equation) is integrated along the streamline
of the self-similar wind solution, recording Γ(θ), the radiative
acceleration. Γ(θ) is then input into the self-similar model in
the subsequent iteration (see Appendix A; specifically, eqn.
A3).
In its simplest form, Euler's Equation is given by
ρ(cid:18) ∂v
∂t
+ (v · ∇)v(cid:19) =XFi,
where the Fi represent the various forces included.
(6)
Radiative Acceleration of MHD Winds
7
the gravitational term can be written as
GM
R2
[1 - Γ(θ)]
R · s = - [1 - Γ(θ)]
r2 + z2 cos(θ - θF). (11)
Next, we take the dot product of s with the magnetic terms to
find
GM
(cid:20)-
1
8πρ
∇B2 + 1
4πρ
(B · ∇)B(cid:21) · s = - Bφ
4πρr
∂(rBφ)
∂s
.
(12)
Combining all of those terms, the full Euler Equation is ob-
tained:
vp
∂vp
∂s
- v2
φ
r
sin θF = - [1 - Γ(θ)]
GM
(r2 + z2)
cos(θ - θF)
- Bφ
4πρr
∂(rBφ)
∂s
.
(13)
This equation is still dependent on vφ, however, which can
be eliminated by appealing to the induction equation and Eφ =
0 for such axisymmetric systems (see, e.g., Königl & Pudritz
2000), to yield a relation between vp and vφ:
vφ =
vpBφ
Bp
+ Ωr.
(14)
Substituting this expression into the Euler Equation yields:
vp
∂vp
∂s
- (cid:18) vpBφ
Bp
- [1 - Γ(θ)]
GM
(r2 + z2)
+ Ωr(cid:19)2 sin θF
cos(θ - θF) -
r
=
Bφ
4πρr
∂(rBφ)
∂s
.
(15)
To evaluate the effective optical depth t, an expression is re-
quired for dvR/dR, the spherical radial gradient of the spher-
ical radial velocity. Since the code integrates quantities only
parallel to the flow, approximations to the perpendicular ve-
locity gradients must be used. Assuming that the derivatives
of θ and θF with distance along the streamline are small (ver-
ified a posteriori to be true):
dvR
dR
= R · ∇vR,
= R · ∇[vp cos(θ - θF)],
≈ cos(θ - θF)(cid:18) dvp
≈ cos(θ - θF)
= cos2(θ - θF)
dvp
ds
dvp
ds
.
sin θ + dvp
dz
cosθ(cid:19) ,
dr
(sin θF sin θ + cosθF cos θ),
(16)
(17)
(18)
(19)
(20)
This integration procedure has been tested with radiative
acceleration turned off, where it reproduces the original self-
similar velocity profile to within one part in 105. With the ra-
diative acceleration turned on, the entire code has repeatedly
converged within approximately eight iterations to an equilib-
rium magnetic wind structure (see Fig. 11). These tests show
that consistent solutions are found and that radiation pressure
does indeed affect the shielded component of the wind.
3.4.1. Critical Points
As with any steady-state wind dynamics problem, one
must search for and consistently pass all critical points (e.g.,
Vlahakis et al. 2000). Critical points mark the location in
the wind where the flow speed is equal to the speed of in-
formation propagation in the wind, and mark locations in
the solutions where solutions branches, or roots, meet; in
the case of radiatively accelerated winds, the location of the
critical point is set by the information propagation speed of
radiative-acoustic wave, or Abbott speed (see, e.g., Abbott
1980; Mihalas & Weibel-Mihalas 1999).
In integrating the
equation of motion for the radiatively-accelerated wind, this
code searches for critical points by looking for multiple roots
in the solution to the equation of motion. However, no critical
points due to radiative acceleration are present in any of the
wind solutions we have found (the magnetocentrifugal wind
does, however, always pass through its own Alfvén critical
point).
To check this result, we have duplicated the work of
Feldmeier & Shlosman (1999, hereafter, FS99), verifying that
for simple wind geometries and without magnetocentrifugal
acceleration, the integration code does indeed encounter a ra-
diative critical point as predicted and found by FS99. In par-
ticular, for the field geometries and forces used in FS99, we
have found identical solutions to both their analytical and nu-
merical calculations.
We then gradually add, to the FS99 model, new components
that are present in our new calculations. When the centrifu-
gal acceleration and the enforced corotation near the base of
the wind are introduced into the framework of FS99, the cen-
trifugal acceleration overwhelms the acceleration of the sec-
ond root that was present in FS99. Therefore, only a single
root is found, and with only a single root, a critical point can-
not be present in our solutions. Another way to check this
is to examine the limit of winds launched at very large an-
gles to the accretion disk. Indeed, for those large angles, the
radiative acceleration begins to dominate the centrifugal ac-
celeration, and the critical point reappears. Therefore, for the
geometry of these magnetocentrifugal winds, where the angle
of the outflow to the accretion disk surface is less than 60◦,
no radiative critical point is expected within these solutions:
a radiative critical point will not be present when centrifugal
acceleration is dominant.
3.5. Model Assumptions and Limitations
This model includes simplifying assumptions about the out-
flow in order to make these calculations possible. In this sub-
section, the assumptions and limitations of the model are sum-
marized, as are the reasons for allowing those assumptions.
First and foremost among these assumptions is self-
similarity. While enabling a relatively quick and flexible
model that can be used to survey a wide variety of outflows,
this assumption does impose constraints on the dynamics of
the model. However, the assumption of self-similarity is es-
sential to the magnetocentrifugal wind solution, as it simpli-
fies the complicated MHD equations. Simultaneously, the ra-
dial self-similarity accommodates, very easily, the radial ge-
ometry for the photoionization simulations.
The above-mentioned Cloudy photoionization simulations
assume a static medium, which is an approximation as well.
As we show in §3.2, however, this approximation is valid for
the these wind models.
Since the accretion disk is a boundary condition in these
models, these winds are assumed to be loaded with matter
from the disk. Accurate models of the accretion disk struc-
ture are beyond the scope of this paper, so it is assumed that
the full mass outflow rate of the the wind is indeed input onto
the magnetic fieldlines at the accretion disk surface. In ad-
-
8
Everett
dition, the matter that is loaded onto those fieldlines is as-
sumed to flow supersonically, i.e., the gas has already passed
the sonic point in the flow. This is done to simplify the mag-
netocentrifugal wind equations and retain the basic model as
outlined in BP82; as such, the same asymptotic expansions
near the disk (from BP82) are utilized here. This treatment
of the wind has been checked in several different ways. First,
the magnetic pressure in the wind is indeed greater than the
thermal pressure throughout the entire wind. Also, the wind's
final velocities are much greater than the sound speed at the
base of the wind, showing that thermal effects are negligible
in determining the final wind velocities. Finally, the lowest
speeds found in the wind model are of order 20% of the sound
speed, and such low Mach numbers are found only very near
the disk, at the base of the wind. Given the above evidence of
the dominance of magnetic fields and radiative acceleration,
the "cold-wind" approximation is valid for these calculations.
Also, in the calculation of the radiative acceleration of the
wind, an approximation to the velocity gradient along spheri-
cal rays is required. This approximation is necessary because
the Euler integration for the wind yields velocity gradients
only along the flow (the poloidal velocity gradients), and thus
the other components must be approximated geometrically
(see §3.4).
The approximations and limitations outlined above do con-
strain the use of this wind model, but in making these com-
promises, a versatile tool can be developed to study the chosen
geometries and forces.
4. RADIATION TRANSFER WITHIN A FIDUCIAL
MAGNETOCENTRIFUGAL WIND
For definitiveness, this model is first employed to examine
the radiative transfer within one fiducial irradiated magneto-
centrifugal wind. For the purpose of this paper, 'fiducial' is
defined to indicate the parameters listed in Table 1. These pa-
rameters are not meant to represent a proposed model for any
one particular AGN, but to define a starting point from which
to examine the structure of these outflows as well as the de-
pendences of the outflows on the model parameters. For in-
stance, the shielding column of NH,shield,0 = 1023 cm- 2 (where
NH,shield,0 represents the value of NH,shield at the base of the
wind, i.e. just above the accretion disk surface) is chosen be-
cause it displays an amount of radiative acceleration between
the extremes of the smaller and larger columns that will be
tested. Similarly, the radiative wind column that is defined is
again between the extremes of nearly optically thick NH,rad,0 =
1023 cm- 2 and very optically thin NH,rad,0 = 1019 cm- 2. Study-
ing such a model first will help bring into focus important
issues concerning the interplay of dynamics and photoioniza-
tion, and represents a foundation from which one can explore
the parameter dependencies of the model.
This fiducial model was therefore run with the parameters
given in Table 1, and after eight iterations, converged to the
final wind structure. The results for the fiducial model are
shown in Figures 4 though 11. In these figures, an overview
of the equilibrium state of this model is presented. The results
displayed in these plots are discussed in detail below.
First, Figure 4 shows the height of the poloidal streamline
as a function of radius in units of the launching radius. In ad-
dition, to illustrate the small difference in geometry between
the final and initial wind models, Figure 5 gives the fractional
change in height as a function of radius. Both of these figures
show that the wind still maintains a collimated state, achiev-
ing a height of z/r0 ∼ 100 (where r0 is the launching radius)
at a cylindrical radius of only r/r0 ∼ 30. So, in this fiducial
model, despite the input from radiative acceleration, the wind
maintains this streamline with only small changes in the struc-
ture of the wind throughout all iterations (see Fig. 5). Thus,
for the case of the fiducial model with L/LEdd = 0.01, this
added acceleration does not significantly affect the structure
of the magnetocentrifugal outflow. These models do show
changes in the velocity structure of the wind near the disk
surface (as will be shown in Fig. 11), but the poloidal wind
structure does not change significantly: on the scale of Fig-
ure 4, the streamlines of the initial, purely magnetocentrifugal
streamline would lie on top of the streamline shown.
FIG. 4. -- Poloidal wind streamlines in the fiducial model. Both the cylin-
drical radial coordinate, r, and the height, z, are given in terms of the launch
radius, r0. Note that the wind is still somewhat collimated, since the radiative
acceleration input is fairly low for L/LEdd = 0.01.
FIG. 5. -- Fractional difference in streamline height as a function of radius
between the final iteration and first iteration of the fiducial model. The small
different in height emphasizes that the structure of the magnetocentrifugal
wind is not significantly modified for the fiducial model's low L/LEdd = 0.01.
Radiative Acceleration of MHD Winds
9
PARAMETERS ADOPTED FOR THE 'FIDUCIAL' MODEL IN THIS STUDY.
TABLE 1
Parameter
Fiducial Value
Parameter Description
109 cm- 3
1023 cm- 2
1021 cm- 2
108M⊙
n0
NH,shield,0
NH,rad,0
M•
Incident Spectrum Risaliti & Elvis (2004)
Lcontinuum
r0
κ
λ
b
Dust in Wind
0.01 LEdd
3 × 1016 cm
0.03
30.0
1.5
No
initial density of the wind at the launch radius
gas shielding column at the base of the wind
gas column, behind the shield, that is radiatively accelerated
mass of the central black hole
Spectrum for the central continuum
luminosity of the central continuum
launch radius of the wind
dimensionless ratio of mass flux to magnetic flux in the wind
normalized total specific angular momentum of the wind
power-law describing variation of density with spherical radius in the wind: n ∝ R- b
presence of dust in the wind
At logarithmically-spaced co-latitudinal angles along the
streamline, Cloudy photoionization simulations are run to de-
termine the photoionization state of the gas, as well as the
radiative transfer through the shield and wind. Changes in
the continuum transmitted through the shield are shown in
Figure 6; the various plots show the simulated continuum at
various heights in the shield corresponding to the indicated
columns. As the shielding column decreases as a function
of height above the disk, the shield transmits progressively
more and more of the ionizing radiation. This plot also dis-
plays how rapidly the column drops as a function of height
above the disk: NH,shield,0 = 1023 cm- 2 occurs at θ = 89.9◦,
NH,shield = 1022.5 cm- 2 at θ = 89.8◦, NH,shield = 1022 cm- 2 at
θ = 89.4◦, and NH,shield drops to 1021 cm- 2 at θ = 85.1◦.
The flux transmitted through the shield then illuminates the
radiatively accelerated wind; results from the photoionization
simulations for the wind are presented in Figure 7, where the
streamline, velocity, density, ionization parameter and tem-
perature in the wind are plotted. In Figure 7a, the height of a
wind streamline as a function of distance along the streamline
(labeled s, given in units of the initial radius, r0) is shown; this
plot simply recasts the structure of the flowline shown in Fig-
ure 4 in terms of s for comparison with the remaining plots.
In Figure 7b, velocities along the streamline are plotted,
showing not only the rapid acceleration in the wind, but also
comparing the components of the wind's velocities. All ve-
locities are given in units of the Keplerian velocity at the base
of the wind, vk,0 (vk,0 = 6.65 × 103 km s- 1 for the parameter
values in Table 1). This plot shows that the vertical velocity is
dominant in these winds at large distances (again showing the
wind is somewhat collimated), with the radial and azimuthal
velocities becoming approximately equal far from the launch-
ing radius of the outflow. Near the very base of the disk, the
radial velocity quickly dominates both the azimuthal and ver-
tical velocities (qualitatively similar to the velocity structure
calculated for radiatively-dominated flows as in MCGV95).
Most importantly, though, we note the extraordinarily rapid
acceleration of the gas from the disk; such acceleration is a
hallmark of both magnetocentrifugal as well as radiatively-
dominated winds, which usually accelerate to their terminal
velocities in a distance on the order of their launching radius.
Due to mass conservation, the extremely rapid accelera-
tion of the wind causes a sharp drop in both the number den-
sity and the column density with height above the disk: both
the number density and column density immediately drop by
three orders of magnitude as the wind rises above the disk.
This is displayed in Figure 7c. This overall drop in density
is extremely important for the ionization state of the wind,
not only for the observational ramifications (i.e., what ions
are present in various parts of the wind) but for the accel-
eration of the wind as well, as will be shown very shortly.
Figures 8 and 9 show in more detail how the radiative accel-
eration leads to a substantial change (by approximately a fac-
tor of two) in both velocity and density with height near the
disk surface. The changes in velocity for the pure magneto-
centrifugal wind as compared to the magnetocentrifugal and
radiatively-accelerated wind is shown in Figure 8. The dif-
ference in the density profile for the pure magnetocentrifugal
wind as compared to the magnetocentrifugal and radiatively-
accelerated wind is shown in Figure 9. (Note that all densities
are normalized to their value at the base of the wind.)
The corresponding ionization state of the wind and temper-
ature are shown in Figure 7d. Most striking is the dramatic
rise in the ionization parameter as the wind rises above the
disk, which is simply due to the drop in density and in shield-
ing already mentioned. The ionization parameter is of prime
interest, as the radiative acceleration in resonant lines is de-
pendent on the number of atomic lines in the gas; the rapid
ionization of the gas prompts questions about how efficient
line-driving will be within this magnetocentrifugal wind. In
addition, can a wind with such a dramatic drop in column
density form an effective "shield"? And for these models,
how does the radiative acceleration then compare to magnetic
acceleration?
These questions are addressed in Figure 10. From the
Cloudy simulations summarized in Figure 7, both the bound-
free and bound-bound radiative acceleration are calculated;
the resultant acceleration (compared to the local gravity) for
the fiducial model is shown in Figure 10. This model shows
that both line-driving and continuum-driving have important
roles to play, with the line-driving dominating the contin-
uum driving in the high-density, low-ionization part of the
wind, and continuum driving dominating line-driving at larger
distances, when the density is much lower.
It is important
to note that for the parameters of the fiducial model with
L/LEdd = 0.01, magnetic acceleration is still much greater than
either line or continuum acceleration. Line-driving is greater
than continuum-driving in only part of the outflow (although
this can change with the density at the base of the wind, as will
be shown in §5.2), as the ionization parameter is low enough
only at the base of the wind for significant numbers of atomic
lines to exist.
In addition, radiative driving is not immedi-
ately important at the extreme base of the outflow, because of
10
Everett
FIG. 6. -- Both the incident continuum on the shield and the transmitted continuum through the shield is displayed. Note the rapid decline in the shield's
column density as a function of polar angle: NH,shield,0 = 1023 cm- 2 lies at the polar angle θ = 89.9◦, NH,shield = 1022.5 cm- 2 at θ = 89.8◦, NH,shield = 1022 cm- 2 at
θ = 89.4◦; the radiation encounters NH,shield 1021 cm- 2 at only θ = 85.1◦. Also indicated for each column is the distance along the flowline (s) where that column
occurs, given both in cm and in units of the launching radius, r0.
the low fluxes that penetrate the columns there. Meanwhile,
the acceleration due to continuum-driving stays close to the
Eddington ratio, at ∼ 0.01. This value is reasonable, as most
of the continuum acceleration comes from electron scattering,
so that Γ ∼ 0.01 would be expected. Minor increases above
that value very near the disk surface are due to bound-free
transitions in the portion of the wind closer to the disk, where
ΓContinuum rises to ∼ 0.02.
As has already been shown, at this low Eddington ratio the
structure of the wind does not change significantly. However,
the velocity at the base of the outflow is affected. The change
in velocity due to radiative acceleration is shown in Figure 11.
This figure shows both the poloidal velocity as a function of
distance along the flowline, and the variation in that velocity
with iterations of the model, therefore showing the conver-
gence in the model. Figure 11 shows that line-driving near
the disk does significantly accelerate the wind, but magnetic
driving determines the terminal velocity at larger radii. This
figure also displays how the code converges; the relatively
slow convergence during the first four iterations is the result of
the program slowly increasing the radiative acceleration to the
computed value (increased slowly in order to avoid severely
overestimating the radiative acceleration in the lines and caus-
ing sudden deceleration in later iterations). In the later itera-
tions, the calculation converges to a final velocity profile using
the full radiative acceleration. This profile shows the affect
of line-driving near the base of the wind, where the velocity
increases above that of the initial magnetocentrifugal wind.
However, as already mentioned, the magnetocentrifugal wind
(in this model, where L/LEdd = 0.01) still determines the ve-
locity at large distances. At those distances, the gas is too
ionized to be appreciably accelerated by line driving.
It is important to note that considering non-Sobolev effects
leads to large changes in Mlines, the line force multiplier, found
in the above calculations: "capping" the Sobolev absorption
length-scale by the actual absorbing column length leads to
smaller optical depths and larger line accelerations. This is
illustrated in Figure 12, where the fiducial model has been
calculated with the Sobolev approximation as well as our non-
Sobolev treatment (see Section 3.3.2). The difference in force
multiplier is due to the strict Sobolev treatment overestimat-
ing the column for the low-acceleration gas near the base of
the wind. This relatively straightforward modification is very
important to correctly estimate the optical depth, as can be
seen in Figure 12.
Overall, in this section, the fiducial model has shown how
the wind velocity, number density, column density, and radia-
tive acceleration all interact to determine the final state of a
magnetocentrifugal wind. These components have not previ-
ously been self-consistently combined in a magnetocentrifu-
gal model, and so yield a new look at the state of these winds.
In addition, the importance of the number densities and col-
umn densities to the final result are most apparent, and clearly
merit further investigation, which will be addressed in §5.
5. DEPENDENCE OF WIND STRUCTURE ON MODEL PARAMETERS
Having analyzed the fiducial model in detail, and observed
how that model's properties change with height, how sensitive
are the trends in §4 to those fiducial parameters? This is an
Radiative Acceleration of MHD Winds
11
FIG. 7. -- The geometry, dynamics, and photoionization state of an illuminated, radiatively-accelerated magnetocentrifugal wind. Pane (a) shows the height
of the wind (in units of the launching radius) with respect to distance along the flowline, s. The arrows in pane (a) give the positions of the columns shown in
Figure 6 (1023, 1022.5, 1022, 1021 cm- 2; decreasing with increasing s and z). Pane (b) shows the various components of the velocity: poloidal, radial, azimuthal,
an vertical, respectively, in units of the Keplerian speed at the base of the flow (for the fiducial model, that base Keplerian speed, vk,0, is 6.65 × 103 km s- 1). Pane
(c) displays various densities: the total hydrogen and electron number densities as well as the total hydrogen column and the effective Hydrogen column (defined
as the column of cold, neutral gas of solar abundances that would produce the same obscuration at 1 keV). All densities are plotted in units of the number density
or total column density at the base of the wind. Pane (d) shows both the electron temperature in the gas and the ionization parameter, U = nγ /nH where nγ is the
number density of hydrogen-ionizing photons.
important question, and one of the key attributes of this self-
similar model is that it allows some flexibility in the selection
of initial parameters. In this section, we test for variations in
the wind by examining how the wind changes as parameters
are modified.
5.1. Variations with Shielding Column
One of the most difficult issues for radiative driving in
AGNs is over-ionization of the wind. As shown in §4, as the
wind accelerates and its density decreases, the magnetocen-
trifugal outflow can easily become too ionized to be efficiently
accelerated to escape velocity solely by atomic lines. This is
the problem of the "shielding gas" that was mentioned in §1:
for pure radiative line-driving, some shielding gas is required
to intercept the X-ray ionizing radiation so that the remaining
UV resonant line photons can be absorbed by the wind and
radiatively accelerate it to the escape velocity.
Some important papers have already been dedicated
to examining the concept of
such as
Chelouche & Netzer (2003b, considering very detailed pho-
toionization simulations of gas shields with constant col-
umn density) and Proga & Kallman (2004, where multidi-
shielding gas,
mensional hydrodynamics simulations with approximate ra-
diative effects are considered). In contrast, the models pre-
sented here include a shield where the column density varies
with height in a shielding wind, and where detailed photoion-
ization simulations can be employed.
Is it
The MHD wind model presented here already launches
is immune to con-
a wind magnetocentrifugally, so it
cerns of overionization.
therefore possible for a
magnetocentrifugally-driven wind, with its commensurate
drop in column density with height above the disk, to act as a
shield, allowing for more efficient radiative acceleration be-
yond it? We can test this question by simply varying the
shielding column in the fiducial model and checking the radia-
tive acceleration seen by a wind launched behind the shield.
As shown in Figure 13, as the shielding column is in-
creased from NH,shield,0 ∼ 1021 to 1024 cm- 2, line-driving in
the wind increases from Γlines ∼ 0.05 to Γlines ∼ 0.12. (Recall
that for this model, continuum driving is Γcontinuum ∼ 0.01.)
This shows that, for large shielding columns (NH,shield,0 ∼
1024 cm- 2), line driving can be up to an order of magnitude
more effective than continuum-driving at the base of the wind.
This increase in acceleration is due to the absorption of the
12
Everett
FIG. 8. -- Variation in all components of the velocity between the initial, pure magnetocentrifugal wind model and the final magnetocentrifugal & radiatively-
driven model. The poloidal velocity (shown in pane a) shows the effects of radiative acceleration (increasing the velocities near the base of the wind by
approximately a factor of two for this low L/LEdd = 0.01), as do the poloidal velocity's components, the radial and vertical velocities (panes b and d). The
azimuthal component of the velocity is relatively unaffected (pane c).
ionizing radiation by the shield, which allows a lower ion-
ization state in the wind, and more line-driving due to more
atomic lines. It is also apparent that, as the shielding is in-
creased, the resultant lower total flux at the base of the wind
means that the onset of significant radiative acceleration is
delayed: this accounts for the offset of maximum radiative
acceleration from the disk surface as the shielding column is
increased.
Further, in Figure 14, the ratio of radiative acceleration to
magnetic acceleration along a streamline is shown. Since the
MHD effects are still supplying most of the acceleration (with
an acceleration roughly equal to and opposite that of gravity),
the ratio of radiative to magnetic acceleration looks much like
that in Figure 13. Magnetic effects dominate in these wind
models, even when large columns of shielding are included.
We have therefore shown that a magnetocentrifugal outflow
can act as a shield and increase the efficiency of line-driving
in the wind. However, it can also be seen that line-driving is
important in these models only at the base of the wind. This
arises not only from the drop in the shield's column density
with height above the disk, but the drop in the wind's density
as well (and the commensurate rise of the ionization parame-
ter).
5.2. Variations with Initial Density
Owing to the increase of line-driving with decreasing ion-
ization parameter, higher accelerations would also be ex-
pected at higher densities. Thus, we investigate the effect of
changes in the initial density in the wind in Figures 15 and 16.
Displaying the effect of a range of initial densities, Fig-
ure 15 shows that line-driving is only effective in these mag-
netocentrifugal winds at relatively high densities. Since con-
tinuum driving is approximately constant (and relatively inde-
pendent of density) at a/g ∼ 0.01 for L/LEdd = 0.01, any line-
driving below that level is insignificant for these winds. Thus,
for initial densities n0 < 109 cm- 3, line driving falls below the
level of continuum driving and ceases to be important. For the
highest density tested, n0 = 1011 cm- 3, line driving dominates
continuum driving for all locations in the wind. (The varia-
tions in each acceleration curve as a function of s shown on
this plot are chiefly due to variations in ionization parameter:
line-driving is high near the disk due to shielding and the rela-
tively high density, and rises towards the end of the streamline
due to the dropping flux levels at large distances.)
Since n0 clearly has a great impact on the radiative acceler-
ation, how do such changes affect observables, such as the ve-
locity? Figure 16 shows how the variation in radiative accel-
eration affects the poloidal velocity (vp) of the outflow. As the
initial density and radiative acceleration increase, the wind's
velocity shows substantial variations from the pure magneto-
Radiative Acceleration of MHD Winds
13
FIG. 9. -- Variation in density between the initial, pure magnetocentrifu-
gal wind model and the final magnetocentrifugal & radiatively-driven model.
Shown here is the change in the hydrogen density with height, but the elec-
tron density and columns change similarly: the added radiative acceleration
yields a drop in density near the base of the disk, as can be seen in all of the
various density measurements.
FIG. 11. -- Poloidal velocity for the fiducial wind as a function of dis-
tance along the flowlines, showing the evolution of the velocity over several
iterations of the code. Line driving does significantly affect the magneto-
centrifugal wind near the base, but magnetocentrifugal driving dominates at
larger distances for this Eddington ratio. (In the first iteration in this plot,
only magnetocentrifugal driving has been considered, so the "Iteration #1"
curve shows the pure magnetocentrifugal wind case.)
FIG. 10. -- Line radiative acceleration, continuum radiative acceleration,
and magnetocentrifugal acceleration along the gas streamlines, compared in
the fiducial model. Note that at the high densities and columns near the base
of the wind, line-driving dominates continuum driving, but that both are less
than the magnetic driving for a system at L/LEdd = 0.01. At larger distances,
continuum driving dominates line driving in the more highly ionized gas. As
expected, continuum-driving is approximately of order the Eddington ratio
for the highly-ionized portion of the outflow.
centrifugal model (which dominates the n0 = 107 cm- 3 and
n0 = 108 cm- 3 models). In the case of n0 = 1011 cm- 3, where
the greatest difference in vp is seen, the velocity increases
by a factor of ∼ 2 to 3 close to the disk. Beyond the re-
gion close to the disk (s/r0 > 1), however, magnetocentrifu-
gal driving still dominates the final velocities for these winds.
But the velocity differences near the disk may be observa-
tionally important, especially if acceleration near the base of
the wind is the source of single-peaked emission lines (as in
Murray & Chiang 1997). If true, such emission lines may be
critical in testing the differences between wind models.
5.3. Variations with Radiative Column
FIG. 12. -- Comparison of the line force multiplier, MLines (see Eq. 4),
for the case of our non-Sobolev method (defined in Section 3.3.2) and the
strict Sobolev calculation. The over-estimate of the absorbing column in the
Sobolev approximation leads to higher opacity in the gas (higher t) and thus
lower acceleration.
Having already tested the more obvious parameters of the
initial density and shield column density, we now turn to
one of the most crucial parameters for the efficiency of line-
driving: the optical depth in the lines.
Under normal circumstances, where the gas velocity is suf-
ficiently low, or where the gas column is very low, the optical
depths in the lines are simply governed by the ionic columns
themselves. For large accelerations or large columns, how-
ever, the optical depths are dominated by the Sobolev length,
which is defined as the distance over which the relative ve-
locity between atoms is equal to the thermal width, so that
a photon emitted by one atom could be absorbed by another
within that Sobolev length (see §3.3.1). With the Eddington
ratio in this magnetocentrifugal wind model (L/LEdd = 0.01),
14
Everett
FIG. 13. -- Variations in line-driving as the magnetocentrifugally-launched
shielding column is increased. With greater shielding columns, the wind is
less highly ionized, leading to higher radiative acceleration near the base of
the wind. Also, with increased shielding comes lower flux levels at the base
of the wind, which displaces the onset of line-driving to larger distances from
the launch point.
(Continuum-driving is approximately constant at a/g ∼
0.01.)
FIG. 15. -- Variations in line-driving with changes in the density at the
base of the wind. Higher densities lead to smaller ionization parameters and
much larger acceleration. (Continuum-driving is approximately constant at
a/g ∼ 0.01.)
FIG. 14. -- As in Figure 13, but comparing the strength of the total radiative
acceleration (both line and continuum acceleration) to the acceleration due to
magnetic fields, where both are normalized relative to gravity. To compare
with the magnetic field acceleration, the radiative acceleration is scaled by a
geometrical factor (see Eq. 13) that yields the radiative acceleration along the
streamline.
both regimes can be important. Depending on the initial pa-
rameters prescribed, the wind column can be small enough
such that the column alone determines the opacity in the wind
(instead of the velocity) and therefore the amount of observed
acceleration, so that the Sobolev length is not important. This
is demonstrated in Figure 17, where the variation in line-
driving with the radiatively accelerated wind column is pre-
sented (the shielding column is held constant). For the larger
columns, the optical depth in the lines increases and the ra-
diative acceleration decreases. For the smallest columns, very
large radiative acceleration is predicted due to the low opac-
FIG. 16. -- As in Fig. 15, except that the change in velocity along the
streamline is shown for the variety of densities. For this Eddington ratio,
the magnetocentrifugal wind still dominates, but as the density increases, the
importance radiative driving increases.
ity in the lines. This is critical for these models, for at sig-
nificantly high column, the wind would see no significant line
driving (this is true for magnetocentrifugal wind columns with
NH,rad,0 & 1022 cm- 2).
5.4. Modifying the Eddington Ratio
Of key importance to applications to AGN is understanding
the acceleration of outflows as a function of the Eddington
ratio, L/LEdd. To investigate the impact of varying Eddington
ratios in our model, we present Figure 18, which displays the
radiative acceleration (both the combined line and continuum
acceleration in panel a as well as the line driving in panel
b) in the fiducial model for three different Eddington ratios:
L/LEdd = 0.001,0.01, and 0.1.
The largest variation in Figure 18a is the continuum driving
Radiative Acceleration of MHD Winds
15
of approximately two to three NH,shield,0 = 1023 cm- 2
(§5.1) and by up to almost two orders of magnitude
for NH,shield,0 = 1024 cm- 2 (§5.4). A magnetocentrifugal
wind has the advantage that it can be accelerated with-
out regard to the ionization state, whereas radiatively-
driven winds must have a low ionization parameter in
order for a critical abundance of atomic lines to be
present. Therefore, magnetocentrifugal winds could
play an important role in acting as a radiation shield
and allow large radiative accelerations. It may also be
possible that pressure differences (MCGV95) or disk
photons (Proga & Kallman 2004) may help "lift" the
shield; neither of those effects are considered here.
Later work with this model will include the effect of
disk-emitted photons.
FIG. 17. -- Variations in line-driving as the radiative wind column is in-
creased. As the wind thickness increases, the optical depth within of the
lines within the magnetocentrifugal wind increases and line-driving drops in
strength. (Continuum-driving is approximately constant at a/g ∼ 0.01.)
increasing linearly with the Eddington ratio. As expected, the
continuum acceleration, relative to gravity, is roughly equal to
the Eddington ratio. The line driving, on the other hand, can
be seen in both the deviations from the approximately con-
stant continuum acceleration in Figure 18a and in Figure 18b.
We begin examining this figure by concentrating on the first
three models in Figure 18, which have NH,shield,0 = 1023 cm- 2.
In these models in Figure 18a, the increase in acceleration due
to line-driving, relative to the continuum-driving, decreases
as the Eddington ratio increases. This can also be seen in
Figure 18b, where the line-driving peaks near the disk (s/r0 .
0.01) for L/LEdd = 0.01, but decreases for Eddington ratios
an order of magnitude larger (where the gas is overionized)
and an order of magnitude smaller (where the radiation field
doesn't have the momentum to accelerate the wind as strongly
as at L/LEdd = 0.01).
Now we turn to the fourth model in Figure 18, where
we keep L/LEdd = 0.1 but increase the shielding level to
NH,shield,0 = 1024 cm- 2. This model shows the importance of
shielding gas for L/LEdd = 0.1. The increase in L/LEdd and
the increase in shielding relative to the fiducial model allow
for increased line radiative acceleration that jumps almost two
orders of magnitude in strength near the base of the wind.
6. RESULTS
If magnetocentrifugal winds power outflows in AGNs, they
must certainly be affected by the intense radiation field they
experience; in turn, such winds will also influence the ef-
ficiency of radiative acceleration. This paper has explored
the radiative transfer through these magnetocentrifugal winds,
and how they both are affected by and affect radiative accel-
eration of outflows from AGNs. The model has been used
to explore the detailed dynamics and ionization of a fiducial
magnetocentrifugal disk wind, showing the inter-relation be-
tween shielding column, initial number density, outflow ve-
locity, Eddington ratio, and acceleration.
As a result of this study, these models have shown:
2. The efficiency of line-driving is strongly dependent on
the density at the base of the wind (§5.2). This is due to
the very critical dependence of line acceleration on the
ionization parameter. The lower the ionization state,
the more lines exist to aid in the momentum transfer
from outward-streaming photons. The density at the
base of the disk is therefore crucial to setting to line-
driving within these magnetocentrifugal models.
3. Small columns (NH,rad,0 . 1021 cm- 2) within magneto-
centrifugal winds can be significantly accelerated by
line-driving (§5.3). This point demonstrates the im-
portance of "non-Sobolev" effects;
that at low
columns, the optical depths in the lines drop below the
opacity given by the Sobolev length, and at such low
columns, the radiative acceleration can be underesti-
mated by the simple Sobolev approximation.
i.e.,
In addition, by examining the fiducial model and the above
cases where model parameters are varied, these solutions have
displayed the importance of considering the detailed interac-
tion between the dynamics and photoionization in AGN out-
flows. Calculations of the ionization parameter along the flow
and in the variation of shielding and optical depth along the
flow are central issues to modeling these winds. The issues
outlined above are a few of the dependences that arise from
such modeling, and may indeed (by the variation in accelera-
tion and therefore velocity along the streamlines) lead to tests
to observationally determine the physics of wind launching in
AGNs.
Thus, while this model has been developed to help address
the above questions about shielding and the affects of radia-
tive driving on magnetocentrifugal winds, the solutions avail-
able are not limited to the above, examined cases. Future
papers will study further the variation of radiative accelera-
tion on model parameters such as the SED, atomic line lists
used to calculate the acceleration, initial densities, Edding-
ton ratios, and other parameters of the model. The model can
also be used to explore the absorption and emission features
from such a wind, as well as, for instance, the possible role
of "clouds" within a continuous wind (Everett et al. 2002). In
addition, this model is in no way constrained to only study
AGNs. The same basic physical framework could also be
employed to study winds from accretion disks surrounding
young stellar objects or cataclysmic variables.
7. ACKNOWLEDGMENTS
1. A magnetocentrifugal outflow, acting as a "shield",
can improve the efficiency of line-driving by factors
I gratefully acknowledge my advisor, Arieh Königl, for
many useful conversations throughout the course of this
16
Everett
FIG. 18. -- Variation in both continuum and line driving (a) and line driving alone (b) when the Eddington ratio in the fiducial model is modified. The first
three models all have NH,shield,0 = 1023 cm- 2, while the last model has NH,shield,0 = 1024 cm- 2. For lower Eddington ratios (L/LEdd = 0.001, 0.01) the gas is in
a relatively low ionization state for line-driving to be important, and the acceleration departs from the pure continuum acceleration (which sets the acceleration
relative to gravity to about the Eddington ratio). For a high Eddington ratio (L/LEdd = 0.1), the gas is overionized and only continuum acceleration is important;
raising the shielding to a nearly Compton-thick NH,shield,0 = 1024 cm- 2 lowers the ionization state of the gas and allows much higher radiative acceleration.
project. In addition, many thanks to Gary Ferland and his col-
laborators for developing Cloudy, making it freely available,
and supporting it. The referee's comments were very help-
ful, and those comments helped improve the paper. Thanks
also to David Ballantyne, Pat Hall, Lewis Hobbs, John Kartje,
Ruben Krasnopolsky, Bob Rosner, Nektarios Vlahakis, and
Don York for valuable comments. This work would not have
been possible without the generous support of NASA's ATP
program, in this case via grant NAG5-9063, and the support
of the Natural Sciences and Engineering Research Council of
Canada. This research has made use of NASA's Astrophysics
Data System.
APPENDIX A: DERIVATION OF THE SELF-SIMILAR CENTRIFUGAL WIND EQUATIONS
APPENDIX
In this appendix, a rederivation of the system of self-similar wind equations for the magnetocentrifugal wind is presented. The
equations utilized in this calculation advance upon those presented in BP82 and KK94: the wind not only has an arbitrary density
power-law index, b, as in KK94, but energy conservation is not required. Since the radiation field continually inputs energy into
the outflow, this is an important modification that was not fully considered in the derivation presented in KK94.
First, a stationary, axisymmetric, ideal, cold MHD flow in cylindrical coordinates (r, φ,z) is assumed. The equations are based
on both the radial and vertical momentum equations:
- v2
φ
r
∂vr
∂r
+ vz
vr
= - ρ
∂vr
∂z
ρ(v · ∇)vz = - ρ
∂Φ
∂r
∂Φ
∂z
4π (cid:18) ∂Bz
- Bz
∂r
∂B2
+ 1
4π
∂z
1
8π
∂z (cid:19) - Bφ
- ∂Br
4πr
(B · ∇)Bz,
∂(rBφ)
∂r
(A1)
(A2)
where v is the fluid velocity and B is the magnetic field. The thermal term is neglected in the limit that thermal affects are much
less important than magnetocentrifugal and radiative-driving effects. Φ is the effective gravitational potential, defined as
where M• is the mass of the central black hole, and Γ(θ) gives the local radiative-to-gravitational radial acceleration (see eq. [1]).
These equations are solved by first stipulating mass conservation,
Φ = - [1 - Γ(θ)]
GM•
(r2 + z2)1/2 ,
(A3)
and then relating the flow velocity to the magnetic field via
∇ · (ρv) = 0,
v(r) =
kB(r)
4πρ(r)
+ ω(r) × r,
(A4)
(A5)
(e.g., Chandrasekhar 1956; Mestel 1961), where k/4π is the ratio of mass flux to magnetic flux, and ω(r) and ρ(r) are the field
angular velocity and gas mass density of the flow, respectively. Both ω and k are constant along magnetic fieldlines. In addition,
while the specific energy is not constant, the total specific angular momentum
l = rvφ -
rBφ
k
(A6)
is conserved.
-
Radiative Acceleration of MHD Winds
17
Self-similarity is then imposed on this system by specifying
(A7)
(A8)
where vk,0 is the Keplerian speed at the base of the outflow, vk,0 = (GM•/r0)1/2, and the prime indicates differentiation with
respect to χ. At the same time, the above constants are re-expressed in dimensionless form:
r = [r0ξ(χ), φ,r0χ],
v = [ξ ′(χ) f (χ),g(χ), f (χ)]vk,0,
l
λ ≡
κ ≡
(GM•r0)1/2 ,
k(1 + ξ ′2
0)1/2
Bp,0
vk,0,
where Bp,0 is the poloidal magnetic field strength at the base of the wind.
As in KK94, a general power-law scaling of the density and magnetic field along the disk's surface is defined:
ρ0 ∝ r- b
0 ,
- (b+1)/2
B0 ∝ r
0
.
With this self-similar specification, the radial and vertical momentum equations become, after some simplification:
f ξ ′m′
κξJ
f 2ξ ′
ξJ
+ ξ ′′ f 2 -
(λm - ξ2)2
ξ3(m - 1)2 = - ξ[1 - Γ(θ)]S3 -
f
κξJ
(m′ -
f κξ ′J + f κξχξ ′′) = - [1 - Γ(θ)]χS3 + f ξ ′
f
κξJ2 (cid:18)- (1 + ξ ′2)(b + 1)
JS2(cid:19) - κ f
2
(λ - ξ2)
(m - 1)
ξ ′′
ξ
+
ξ
(χ + ξξ ′)ξ ′
(cid:20) (λ - ξ2)
(m - 1)
χ(cid:18) 2ξξ ′
(m - 1)
(- b + 1)
+
2
+ (λ - ξ2)m′
(m - 1)2 (cid:19)(cid:21) ,
κξJ2 (cid:18)- (1 + ξ ′2)(b + 1)
2
+
ξ
J
(cid:19) -
- ξ ′′(χ2 + ξ2)
(χ + ξξ ′)ξ ′
ξ ′κ f (λ - ξ2)(cid:18) (b + 1)(λ - ξ2) - 2(λ + ξ2)
(λ - ξ2)2m′κ f
(m - 1)3
2ξ(m - 1)2
,
(cid:19) +
(A9)
(A10)
(A11)
(A12)
(A13)
(A14)
(A15)
(A16)
(A17)
(A21)
where
m ≡
κ ≡
4πρv2
p
B2
p
k(1 - ξ ′2
0) 1
Bp,0
l
2 vk,0
2
λ ≡
(GMr0) 1
J ≡ ξ - χξ ′,
S ≡ 1/pξ2 + χ2.
= κξ f J = square of poloidal Alfv´en Mach number,
= dimensionless ratio of mass flux to magnetic flux,
= normalized angular momentum,
(A18)
(A19)
(A20)
The two equations (A13) and (A14) define the differential equations for m′ and ξ ′′, which are, respectively, the spatial gradient
in the poloidal Alfvén mach number (gradient with respect to height, χ) and the (cylindrical) radial velocity gradient (again with
respect to χ).
One can see from close inspection of the above equations that many of the terms have a denominator of (m - 1), showing that
when the gas crosses the Alfvén point (where m = 1), the equations become singular. Rewriting and solving the m′ equation for
the value of m′ at the Alfvén singular point:
A = 2ξJ[- 8χκ2λm′ξ ′J3 + 4(1 + b)κ2λξ ′2J2(χ + ξξ ′) + m′2(χ + ξξ ′)(- 2κ2λ2S +
m′
3
2 (λ - S)ξ ′ + ((1 + b) + 2χ2κ2λ(λ - S))ξ ′2 + 2κ2λΓ(θ)J2)]/
(1 + b) + 2κ2λ3 - 4χκ2λ
(cid:20)4ξJ(cid:18) 4κ2λξ ′2J2
S2
+ m′2(χ + ξξ ′)2(cid:19)(cid:21) .
-
-
18
Everett
This constraint is used to start the integral at the Alfvén point with the value of m′
A given by equation (A21).
As covered in the main text, these equations are solved using a "shooting algorithm," integrating from both the Alfvén point
and the disk surface towards an intermediate point. Matching the integrals of three first-order equations (given by the first-order
equation for m′ and the second-order equation for ξ ′′ in eqs. A13 and A14) at the common point allows us to solve for the three
free parameters in the system: ξ ′
0, ξ ′
A, and χA.
REFERENCES
Abbott, D.C. 1980, ApJ, 242, 1183
Antonucci, R. 1993, ARA&A, 31, 473
Arav, N., Li, Z., & Begelman, M.C. 1994, ApJ, 432, 62
Arav, N., Korista, K.T., Barlow, T.A., Begelman, M.C. 1995, Nature, 376,
Arav. N. 1996, ApJ, 465, 617
Arav, N., Barlow, T.A., Laor, A., Sargent, W.L.W, & Blandford, R.D. 1998,
MNRAS, 297, 990
Arnaud, M. & Raymond, J. 1992, ApJ, 398, 394
Baldwin, J., Ferland, G.J., Martin, P.G., Corbin, M., Cota, S., Peterson, B.M.,
388, 350
& Slettebak, A. 1991, ApJ, 374, 580
Baldwin, J., Ferland, G.J., Korista, K., Verner, D. 1995, ApJ, 455, 119
Balsara, D. & Krolik, J.H. 1993, ApJ, 402, 109
Blandford, R.D., & Königl, A. 1979, Astrophys. Lett., 20, 15
Blandford, R.D., & Payne, D.G. 1982, MNRAS, 199, 883 (BP82)
Blandford, R.D. 2001, in ASP Conf. Ser. 224, Probing the Physics of Active
Galactic Nuclei by Multiwavelength Monitoring, ed. B. M. Peterson, R. S.
Polidan & R. W. Pogge (San Francisco: ASP), 499
Bottorff, M., Korista, K.T., Shlosman, I., & Blandford, R.D. 1997, ApJ, 479,
576
200
Krasnopolsky, R., Li, Z.-Y.,Blandford, R. 1999, ApJ, 526, 631
Kuncic, Z., Celotti, A., Rees, M.J. 1997, MNRAS, 284, 717
Laor, A., & Brandt, W.N. 2002, ApJ, 569, 641
Martin, P.G., & Rouleau, F. 1991, in Extreme Ultraviolet Astronomy, eds,
Malina, R.F., Bowyer S., Pergamon Press, Oxford, 341
Mathews, W.G., & Ferland, G.J. 1987, ApJ, 323, 456 (MF87)
Mathis, J.S., Rumpl, W., & Nordsieck, K.H. 1977, ApJ, 217, 425
Meier, D.L., Edgington, S., Godon, P., Payne, D.G., Lind, K.R. 1997, Nature,
Mestel, L. 1961, MNRAS, 122, 473
Mihalas, D. & Weibel-Mihalas, B. 1999, Foundations of Radiation
Hydrodynamics, (Mineola:Dover Publications, Inc.)
Murray, N., & Chiang, J. 1997, ApJ, 474, 91
Murray, N., Chiang, J., Grossman, S.A., & Voit, G.M. 1995, ApJ,451, 498
Netzer, H. 1990, in Active Galactic Nuclei, Saas-Fee Advanced Course 20
(MCGV95)
(Berlin: Springer-Verlag), 57
Ouyed, R., Pudritz, R.E. 1997, ApJ, 482, 712
Ouyed, R., Pudritz, R.E., Stone, J.M. 1997, Nature, 385, 409
Pereyra, N.A., Owocki, S.P., Hillier, D.J., Turnshek, D.A. 2004, ApJ, 608,
454
Peterson, B.M. 1997, An Introduction to Active Galactic Nuclei, (New
York:Cambridge University Press)
Powell, M.J.D. 1970,
in Numerical Methods for Nonlinear Algebraic
Equations, ed, P. Rabinowitz (New York: Gordon and Breach).
Proga, D. 1999, MNRAS, 304, 938
Proga, D. 2000, ApJ, 538, 684
Proga, D. 2003, ApJ, 585, 406
Proga, D., Stone, J.M., Drew, J.E. 1998, MNRAS, 295, 595
Proga, D., Stone, J.M., & Kallman, T.R. 2000, ApJ, 543, 686, PSK00
Proga, D., & Kallman, T.R, 2004, accepted to ApJ (astro-ph/0408293)
Rees, M.J. 1987, MNRAS, 228, 47P
Reichard, T.A., Richards, G,T., Hall, P.B., Schneider, D.P., Vanden Berk,
D,E., Fan, X., York, D.G.; Knapp, G.R.; Brinkmann, J. 2003, AJ, 126,
2594
Risaliti, G & Elvis, M. 2004, in Supermassive Black Holes in the Distant
Universe, ed. A. J. Barger (Boston: Kluwer)
Romanova, M.M., Ustyugova, G.V., Koldoba, A.V., Chechetkin, V.M.,
Lovelace, R.V.E. 1997, ApJ, 482, 708
Safier, P. 1993, ApJ, 408, 115
Sobolev, V.V. 1958 in Theoretical Astrophysics, ed. V.A. Ambartsumian
(London: Pergamon)
Spruit, H.C. 1996 in Physical processes in Binary Stars, eds. R.A.M.J. Wijers,
M.B. Davies & C.A. Tout, (Dordrecht: Kluwer) (astro-ph/9602022)
Steenbruge, K.C., Kaastra, J.S., de Vries, C.P., Edelson, R. 2003, A&A in
press (astro-ph/0302493)
Stevens, I.R, & Kallman, T.R. 1990, ApJ, 365, 321
Tielens, A.G.G.M., McKee, C.F., Seab, C.G., Hollenbach, D.J. 1994, ApJ,
431, 321
Tran, H.D. 2003, ApJ, 583, 632
Urry, C.M., & Padovani, P. 1995, PASP, 107, 803
Ustyugova, G.V., Koldoba, A.V., Romanova, M.M., Chechetkin, V.M., &
Lovelace, R.V.E. 1995, ApJ, 439, L39
Verner, D.A., & Ferland, G.J. 1996, ApJS, 103, 467
Verner, D.A., Verner, E.M., & Ferland, G.J. 1996, Atomic Data Nucl. Data
Tables, 64, 1
Vitello, P.A.J, & Shlosman, I. 1988, ApJ, 327, 680
Vlahakis, N., Tsinganos, K., Sauty, C., Trussoni, E., 2000, MNRAS, 318, 417
Wardle, M., & Königl, A. 1993, ApJ, 410, 218
Weber, E.J. & Davis, L. Jr. 1967, ApJ, 148, 217
Weymann, R.J., Morris, S.L., Foltz, C.B., Hewett, P.C. 1991, ApJ, 373, 23
Bottorff, M., Korista, K.T., & Shlosman, I. 2000, ApJ, 537, 134
Bottorff, M., & Ferland, G. 2001, ApJ, 549, 118
Cassinelli, J. P. 1979, ARA&A, 17, 275
Castor, J.I., Abbott, D.C., & Klein, R.I. 1976, ApJ, 195, 157
Chandrasekhar, S. 1956, ApJ, 124, 232
Chelouche, D., Netzer, H. 2001, MNRAS, 326, 916
Chelouche, D., Netzer, H. 2003, MNRAS, 344, 223
Chelouche, D., Netzer, H. 2003, MNRAS, 344, 233
Chiang, J. & Blaes, O. 2003, ApJ, 586, 97
Contopoulos, J., & Lovelace, R.V.E. 1994, ApJ, 429, 139
Crenshaw, D.M., Kraemer, S.B., Boggess, A., Maran, S.P., Mushotzky, R.F.,
Wu, C. 1999, ApJ, 516, 750
Crenshaw, D.M., Kraemer, S.B., George, I.M. 2003, ARAA, 41, 117
de Kool, M., & Begelman, M.C. 1995, ApJ, 455, 448
Deluit, S.J. 2004, A&A, 415, 39
Draine B.T., & Lee, H.M. 1984, ApJ, 285, 89
Drew, J.E., & Boksenberg, A. 1984, MNRAS, 211, 813
Elvis, M. 2000, ApJ, 545, 63
Elvis, M., Risaliti, G., & Zamorani, G. 2002, ApJ, 565, L75
Emmering, R.T., Blandford, R.D., & Shlosman, I. 1992, ApJ, 385, 460
Everett, J.E., Königl, A., Arav, N. 2002, ApJ, 569, 671
Fath, E.A. 1909, Lick Obs. Bull., 149, 71
Feldmeier, A. & Shlosman, I. 1999, ApJ, 526, 344
Ferland, G.J., Korista, K.T., Verner, D.A., Ferguson, J.W., Kingdon, J.B., &
Verner, E.M. 1998, PASP, 110, 761
Ferreira, J. 2003 in Star Formation and the Physics of Young Stars, eds. J.
Bouvier and J.-P. Zahn (Les Ulis: EDP Sciences), 229 (astro-ph/0311621)
Friend, D.B. & MacGregor, K.B. 1984, ApJ, 282, 591
Fromerth, M.J & Melia, F. 2001, ApJ, 549, 205
Ganguly, R., Bond, N.A., Charlton, J.C., Eracleous, M., Brandt, W.N., &
Churchill, C.W. 2001, ApJ, 549, 133
Gregori, G., Miniati, F., Ryu, D., & Jones, T.W. 2000, ApJ, 543, 775
Hamann, F.W., Barlow, T.A., Chaffee, F.C., Foltz, C.B., & Weymann, R.J.
Hartley, L.E., Drew, J.E., Long, K.S., Knigge, C., Proga, D. 2002, MNRAS,
2001, ApJ, 550, 142
332, 127
Kartje, J.F. 1995, ApJ, 452, 565
Kartje, J.F., Königl, A., & Elitzur, M. 1999, ApJ, 513, 180
Krolik, J.H. 1999, Active Galactic Nuclei: From the Central Black Hole to
the Galactic Environment (Princeton: Princeton University Press)
Krolik, J.H. & Kriss, G.A. 2001, ApJ, 561, 684
Königl, A., & Kartje, J.F. 1994, ApJ, 434, 446 (KK94)
Königl, A., & Pudritz, R.E. 2000 in Protostars & Planets IV, ed. V. Mannings,
A. P. Boss, S. S. Russell (Tucson: University of Arizona Press), 759
Kraemer, S.B., Crenshaw, D.M., George, I.M., Netzer, H., Turner, T.J., Gabel,
J.R. 2002, ApJ, 577, 98
|
astro-ph/0308540 | 1 | 0308 | 2003-08-29T16:01:51 | Intrinsic and dust-induced polarization in gamma-ray burst afterglows: the case of GRB 021004 | [
"astro-ph"
] | Polarization measurements for the optical counterpart to GRB 021004 are presented and discussed. Our observations were performed with the TNG and the VLT-UT3 (Melipal) during the first and fourth night after the gamma-ray burst discovery. We find robust evidence of temporal evolution of the polarization, which is therefore, at least partially, intrinsic to the optical transient. We do not find convincing evidence of wavelength dependence for the intrinsic polarization of the transient, in agreement with current polarization models for optical afterglows. We discuss the role of dust, both in our galaxy and in the host, in modifying the transmitted polarization vector, showing how a sizable fraction of the observed polarized flux is due to Galactic selective extinction, while it is not possible to single out any clear contribution from dust in the host galaxy. We discuss how our data compare to those obtained by different groups showing that a two-component model is required to describe the complete dataset. This is not surprising given the complex lightcurve of GRB 021004. | astro-ph | astro-ph |
Astronomy & Astrophysics manuscript no. 021004
(DOI: will be inserted by hand later)
November 4, 2018
Intrinsic and dust-induced polarization in gamma-ray burst
afterglows: the case of GRB 021004 ⋆
D. Lazzati1, S. Covino2, S. di Serego Alighieri3, G. Ghisellini2, J. Vernet3, E. Le Floc'h4, D. Fugazza5, S.
Di Tomaso5, D. Malesani6, N. Masetti7, E. Pian8, E. Oliva5, and L. Stella9
1 Institute of Astronomy, University of Cambridge, Madingley Road, CB3 0HA Cambridge, UK.
2 INAF -- Osservatorio Astronomico di Brera, via E. Bianchi 46, 23807 Merate (LC), Italy.
3 INAF -- Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy.
4 Service d'Astrophysique, C.E. Saclay, 91191 Gif -- sur -- Yvette Cedex, France.
5 INAF -- Telescopio Nazionale Galileo, Roque de los Muchachos, P.O. Box 5653, 38700 Santa Cruz de la Palma,
Spain.
6 International School for Advanced Studies (SISSA-ISAS), via Beirut 2-4, 34014 Trieste, Italy.
7 Istituto di Astrofisica Spaziale e Fisica Cosmica, CNR, via Gobetti 101, 40129 Bologna, Italy.
8 INAF -- Osservatorio Astronomico di Trieste, via Tiepolo 11, 34131 Trieste, Italy.
9 INAF -- Osservatorio Astronomico di Roma, via Frascati 33, 00040 Monteporzio Catone (Roma), Italy.
Abstract. Polarization measurements for the optical counterpart to GRB 021004 are presented and discussed. Our
observations were performed with the TNG and the VLT -- UT3 (Melipal) during the first and fourth night after the
gamma-ray burst discovery. We find robust evidence of temporal evolution of the polarization, which is therefore,
at least partially, intrinsic to the optical transient. We do not find convincing evidence of wavelength dependence
for the intrinsic polarization of the transient, in agreement with current polarization models for optical afterglows.
We discuss the role of dust, both in our galaxy and in the host, in modifying the transmitted polarization vector,
showing how a sizable fraction of the observed polarized flux is due to Galactic selective extinction, while it is
not possible to single out any clear contribution from dust in the host galaxy. We discuss how our data compare
to those obtained by different groups showing that a two-component model is required to describe the complete
dataset. This is not surprising given the complex lightcurve of GRB 021004.
Key words. gamma rays: bursts -- polarization -- dust -- radiation mechanisms: non-thermal
1. Introduction
It is now well established that gamma-ray burst (GRB)
optical afterglows (OA) can show some degree of linear
polarization. To date, in five cases a positive detection was
obtained: GRB 990510 (Covino et al. 1999; Wijers et al.
1999), GRB 990712 (Rol et al. 2000), GRB 020405 (Bersier
et al. 2003a; Covino et al. 2003a; Masetti et al. 2003),
GRB 020813 (Barth et al. 2003; Covino et al. 2002a),
and GRB 030329 (Covino et al. 2003; Efimov et al. 2003;
Magalhaes et al. 2003). Usually, the polarized flux is not
large, P <∼ 3%, but in most cases it has been possible to
Send
[email protected]
offprint
requests
to:
D.
Lazzati;
e-mail:
rule out that the observed polarization was induced1 by
dust along the line of sight in our own Galaxy (e.g. Covino
et al. 1999, 2003c). More recently, it was also possible to
exclude, at least for some cases, a major contribution to
the observed polarization level due to the interposition of
dust in the host galaxy. In fact, interstellar polarization is
necessarily associated with reddening, and by modelling
the spectral shape of the OA, the total amount of dust in-
terposed on the line of sight can be constrained, allowing
us to put limits on the dust-induced polarization.
This method is however model dependent, and re-
lies also on the knowledge of the dust properties in the
GRB environment (see e.g. Lazzati et al. 2002a). Given
the present uncertainties, the possibility that for at least
some OA a sizable fraction of the observed polarization
is induced by dust in the GRB environment or in the
⋆ Based on observations made with ESO telescopes at the
Paranal Observatory under programme Id 70.D-0111, on data
from the ESO/ST-ECF Science Archive Facility and on obser-
vations made with the TNG under programme TAC 8 01(47).
1 Here and in the following the only dust induced polariza-
tion that we consider is the one due to the dichroism of the
aligned grains and not the polarization induced by scattering.
2
Lazzati et al.: Polarization of GRB 021004
host galaxy cannot be excluded yet. In principle, there
are at least two safe ways to unambiguously detect in-
trinsic polarization: the first is to perform multiple obser-
vations, looking for temporal variation of the polarization
degree and/or position angle. Such a polarization variabil-
ity would also provide a direct link between the dynamics
of the fireball evolution and the geometry of the emitting
region (Ghisellini & Lazzati 1999, hereafter GL99; Sari
1999; Granot et al. 2002). The second is to study the wave-
length dependence of the polarization, possibly extending
it to the infrared, in an attempt to exclude that it follows
the "Serkowski curve" typical of Milky Way (MW) inter-
stellar dust polarization (Serkowski et al. 1975), bearing
in mind, however, that the OA emission could be intrinsi-
cally polarized in a wavelength dependent way.
Recently, some steps in this direction were performed.
First, for GRB 020813 secure polarization variability was
detected, the degree of polarization decreasing on a day
time scale from P ∼ 2% (Barth et al. 2003) to P ∼
0.8% (Covino et al. 2002a), at a fixed polarization angle.
Moreover, again for GRB 020813 (Barth et al. 2003), and
very recently for GRB 030329 (Covino et al. 2003), spec-
tropolarimetry could be performed; in both cases, small
but significant wavelength dependence was found.
GRB 021004 was localized on 2002 October 4 at
12:06:14 UT by the HETE -- II satellite (Shirasaki et al.
2002). In the FREGATE 8 -- 40 keV and in the WXM 2 --
25 keV bands the burst had a duration of about 100 sec-
onds. It thus belonged to the class of long-duration GRBs.
The optical counterpart was identified less than 10 min
after the burst (Fox 2002) as an R ∼ 15.3 fading ob-
ject at the coordinates α2000 = 00h26m54.s69, δ2000 =
+18◦55′41.′′3. The early detection of the OA and its bright-
ness allowed a dense sampling of the light curve (see e.g.
Lazzati et al. 2002b and references therein), the identifi-
cation of several absorption systems, and of a prominent
emission feature in the optical spectra identified as a Lyα
emission line from the host galaxy at a redshift z = 2.328
(Mirabal et al. 2002, ?; Matheson et al. 2002; Møller et al.
2002; Schaefer et al. 2003).
In the following we will discuss our three polarimetric
observations of GRB 021004: one performed in the near in-
frared (NIR) with the Italian Telescopio Nazionale Galileo
(TNG) at the Canary Islands, and two in the visible band
at the VLT. Furthermore, we analyzed a publicly available
spectropolarimetric dataset retrieved from the ESO VLT
Science Archive. We will then compare these measure-
ments, including the polarimetric observations performed
by Rol et al. (2003), with theoretical models.
Fig. 1. R-band lightcurve of GRB 021004 with the posi-
tion of the polarimetric observations marked. Data are
from Bersier et al (2003a); Fox et al. (2003); Holland et
al. (2003); Pandey et al. (2002); Uemura et al. (2003).
imaging polarimetry mode, and with the grism 300 V in
the spectropolarimetry mode.
2.1. TNG observation
The TNG observation (hereafter run 1) started on October
4.915 (9.86 hours after the GRB trigger) and lasted for
∼ 1.8 hours. The optical transient (OT) J magnitude was
derived by the acquisition frames as J = 17.00 ± 0.05
(hereafter 1-σ errors are reported). The observations were
performed under mediocre seeing conditions (1.5′′) in the
large field mode with a scale of 0.25′′/pixel.
2.2. VLT observations
We analyzed three sets of polarimetric observations of
GRB 021004 carried out by the ESO VLT -- UT3 (Melipal).
2. Observations and analysis
First VLT observation
Observations of GRB 021004 made use of two tele-
scopes. First, the TNG, equipped with the Near Infrared
Camera Spectrometer (NICS) and a J filter,
in the
imaging polarimetry mode; second, the ESO VLT -- UT3
(Melipal), equipped with the Focal Reducer/low disper-
sion Spectrometer (FORS 1) with a Bessel V filter in the
Our first VLT observation (here called run 2) started on
October 5.080 (13.82 hours after the GRB trigger) and
lasted for ∼ 1.6 hours (see Fig. 1). The OT was clearly
detected in the acquisition image with a magnitude V =
19.34 ± 0.02 with respect to the USNO -- A2.0 star reported
by Fox (2002), as calibrated by Henden (2002a, 2002b).
Lazzati et al.: Polarization of GRB 021004
3
Second VLT observation
From the ESO archive we retrieved a public spectropolari-
metric observation (run 3). It started on October 5.247
(17.83 hours after the GRB trigger) and lasted for ∼ 2.0
hours. The spectrum covers the range from 350 nm to
860 nm and the observations were performed with the
300 V grism.
Third VLT observation
Eventually, a last VLT observation (run 4) was performed
starting on October 8.225 (89.3 hours after the GRB trig-
ger), and lasting for ∼ 2.8 hours. The OT magnitude was
V = 20.89 ± 0.03 with respect to the same USNO -- A2.0
star.
Fig. 2. Polarization degree, position angle and normalized
spectrum of the public ESO -- VLT spectropolarimetric ob-
servation (run 3). The polarization degree is roughly con-
stant throughout the optical band at P ∼ 1.9% and the
position angle is also constant at ϑ ∼ 118◦. Observations
were performed with the VLT -- UT3 (Melipal) equipped
with FORS 1 in the spectropolarimetry mode with grism
300 V. Filled symbols correspond to the "clean" part of the
spectrum, while empty symbols are relative to absorbed
portions of the spectrum, or the ones contaminated by
strong sky emission lines.
All VLT observations were
under
good/excellent seeing conditions (0.5′′ − 0.9′′) in standard
resolution mode with a scale of 0.2′′/pixel.
performed
Polarimetric standard stars were also observed. One
polarized, BD-125133,
in order to fix the offset be-
tween the polarization and the instrumental angles, and
three non-polarized, WD 0310 -- 688, WD 1615 -- 154, and
BD+284211, to estimate the degree of spurious polariza-
tion possibly introduced by the instruments. In addition,
we have also analyzed the ESO archive spectropolarimen-
tric data for the NGC 2024 NIR1 polarization standard
star, obtaining in all cases a good Serkowski curve com-
pletely consistent with the available data2.
2.3. Data reduction and analysis
The data reduction was carried out with the Eclipse
package (version 4.3.1; Devillard 1997). After bias sub-
traction, non-uniformities were corrected using flat-fields
obtained without the Wollaston prism. For the IR data
sky flat-fields were applied. The flux of each point source
in the field of view was derived by means of aperture pho-
tometry by the Graphical Astronomy and Image Analysis
(GAIA) tools3 (version 2.6-6).
The general procedure followed for the analysis of
imaging polarimetry observations is indeed extensively
discussed in Covino et al. (1999, 2002b, 2003a) and di
Serego Alighieri (1997), while details about the NICS po-
larimetric capabilities are discussed by Oliva (1997).
The VLT Spectra were extracted with the apall tool
included in the IRAF package (version 2.11), which ex-
tracts the spectrum in a fixed-width window, allowing for
a polynomial evolution of the centroid with wavelength.
Suitable IDL routines were also developed to perform an
independent extraction and check for any possible bias. In
this case the extraction of the 1D spectrum from the 2D
frame was performed by fitting a bell-shaped function to
all the vertical stripes of the 2D spectrum, allowing for a
first order variation of the background and for a 4th de-
gree polynomial variation of the flux centroid and width of
the fitting function. The result of this "fitted" extraction
were not entirely consistent with those of a more standard
sliding window, as discussed below.
The results of the spectropolarimetric observations are
shown in Fig. 2. To all our data we have applied a cor-
rection for the wavelength dependence of the half-wave
plate axis as measured by the FORS team4. The plot is
based on the fitting extraction (hereinafter F) which, as
anticipated, is not fully consistent with the more stan-
dard window extraction (hereinafter W). The F extrac-
tion shown in the figure yields a polarization which is
consistent with being wavelength independent. A simul-
taneous constant fit to the polarization degree and an-
gle returns P = (1.88 ± 0.05)% and ϑ = (118 ± 1)◦
2 http://www.eso.org/instruments/fors1/pola.html
3 http://star-www.dur.ac.uk/∼pdraper/gaia/gaia.html
4 http://www.eso.org/instruments/fors1/
4
Lazzati et al.: Polarization of GRB 021004
with χ2/d.o.f. = 45.8/44 (null probability ∼ 60%). The
bin size has been made large enough to make the po-
larization bias (Wardle & Kronberg 1974) unimportant.
Completely consistent results are obtained with the W
extraction if a small window (6 pixels width) is adopted,
while a wider window (20 pixels) yields different results.
The polarization obtained using the wide W extraction is
wavelength dependent, the polarization decreasing from
∼ 2% at 600 nm to ∼ 1% at 800 nm. The resulting polar-
ization also presents a significant dip at P ∼ 1% around
550 nm. In principle, both methods are prone to inadequa-
cies in extracting the data, and it is not possible to decide
a priori which is the best one. The fit and small W extrac-
tion are in fact sensitive to non centrally symmetric 2D
spectra. The large window, instead, can yield an unbiased
extraction of non-symmetric 2D spectra, but is subject
to a varying background and to contribution from nearby
sources. In an "Ockham razor" approach, we decided to
show the result which requires less parameters to be ex-
plained, namely the constant one, warning however the
reader that according to the large window extraction, the
polarization may be wavelength dependent, being smaller
in the red part of the spectrum.
From an independent analysis of the same data, Wang
et al. (2003) report a marginal evidence of increasing po-
larization at λ <∼ 400 nm, across the rest-frame Lyα ab-
sorption features. They interpret this as a consequence of
absorption of the fireball emission by nearby high veloc-
ity clumps of HI atoms, with a covering factor smaller
than, but comparable to, unity. As can be seen in the up-
per panel of Fig. 2, our binned data do not support this
claim, and the same result is obtained with a binning of
∼ 2 nm, comparable to the one adopted by Wang et al.
(2003). However we find that using smaller wavelength
bins (e.g. 0.52 nm) the polarization, not corrected for the
bias (Wardle & Kronberg 1974), steadily increases from
400 nm down to the shortest available wavelength as ex-
pected for a low sensitivity region at the blue edge of the
spectral range. Therefore this UV rise of the polarization
looks consistent with a bias effect, although the possibil-
ity that some of it might be real cannot be completely
excluded.
We stress that the difference between the various spec-
trum extraction methods discussed here are at the ∼ 1%
level. While being small,
if not negligible, for a stan-
dard spectral analysis, these differences become significant
when weak polarization levels are investigated.
2.4. Other observations
Rol et al. (2003) obtained three independent measure-
ments of the optical polarization of the afterglow of
GRB 021004, three with the Nordic Optical Telescope,
in the R-band, and one with the VLT, in the V -band.
They obtain P = (1.17 ± 0.46)% with position angle
ϑ = (184.2 ± 11.4)◦ at ∆t = 0.366 d after the GRB,
P = (1.73 ± 0.51)% with position angle ϑ = (166.4 ± 8.1)◦
Fig. 3. V -band time-resolved polarimetry of GRB 021004.
Filled circles show the polarization degree (upper panel)
and position angle (lower panel) for our VLT runs (run
2 and run 4) vs. the time elapsed since the GRB trigger.
The diamonds refer to the Rol et al. (2003) measurements,
while the asterisk is obtained by filtering with a Bessel
V transmission profile the spectropolarimetric observation
(run 3). The shaded bands show the MW interstellar po-
larization for field stars. Their vertical width corresponds
to the 1-σ uncertainty.
at ∆t = 0.376 d, P < 1% at ∆t = 0.396 d and P =
(1.29 ± 0.13)% with position angle ϑ = (121.8 ± 2.8)◦ at
∆t = 0.666 d.
3. Results and modelling
3.1. Imaging polarimetry
The IR polarimetric observation (run 1) provided Q and U
Stokes parameters compatible with those derived for run 2
(see below and Tab. 1), even though the larger errors only
allowed us to derive a P < 5% upper limit (95% confi-
dence level). This limit is not particularly stringent com-
pared to those derived in the optical for other afterglows
(e.g. Hjorth et al. 1999; Covino et al. 2002b; Bjornsson et
al. 2002; Covino et al. 2003d). However, it is the first IR
polarimetric observation that provides a useful constraint,
since the previous upper limits were rather loose (∼ 30%;
Klose et al. 2001).
The results of VLT V -band imaging polarimetry, as
well as a synthetic V -band measurement derived from the
spectropolarimetric measurement, are plotted in Fig. 3
as a function of the time elapsed since the GRB trigger.
We also report the value of polarization induced by the
MW selective extinction, as derived by averaging the Q
and U parameters for several bright stars in the field. We
obtain PMW = (0.81 ± 0.03)% and ϑMW = (107 ± 1)◦.
Lazzati et al.: Polarization of GRB 021004
5
Table 1. The normalized polarization Stokes parameters not corrected for the interstellar polarization. Observations
were performed with the TNG (run 1) and with the VLT -- UT3 (runs 2 and 4). VLT -- UT3 spectropolarimetry (run 3)
was performed with grism 300 V. The reported results for the spectropolarimetric observation of run 3 were obtained
by integrating the spectrum over the V -band. Uncertainties are at 1-σ and the upper limit is a 95% confidence level.
Run Tel.
TNG
1
2
VLT
VLT
3
4
VLT
Filter UT (10/02) Magnitude
Q
U
J
V
V
V
4.953
5.172
5.247
8.225
17.00 ± 0.05 −0.0160 ± 0.0130 −0.0222 ± 0.0130
19.34 ± 0.02 −0.0083 ± 0.0009 −0.0094 ± 0.0010
--
20.89 ± 0.03
−0.0075 ± 0.001
−0.0067 ± 0.002
−0.016 ± 0.001
+0.0002 ± 0.002
P (%)
< 5
1.26 ± 0.10
1.74 ± 0.20
0.67 ± 0.23
ϑ (◦)
--
114 ± 2
122 ± 2
89 ± 10
Such a polarization is slightly larger than the maximum
interstellar (ISM) polarization according to the empirical
law PISM <∼ 0.09 EB−V (Serkowski et al. 1975), given the
reddening EB−V ∼ 0.06 estimated for the GRB 021004
field (Schlegel et al. 1998). We note however that devi-
ations from the above law are observed especially along
low-reddening lines of sight (see Fig. 9 of Serkowski et al.
1975). An important conclusion that can be drawn from
Fig. 3 is that neither the polarized fraction nor the posi-
tion angle of the OT were constant during the evolution of
the afterglow (Rol et al. 2003). We conclude that the rel-
ative contribution of the polarizing components, namely
the MW ISM, the GRB host ISM, and the OT itself, did
vary on short time scales. Since the ISM polarization can-
not vary on short (day) time scales, we conclude that the
OT was intrinsically polarized. The rotation of the polar-
ization angle between the first measurements of Rol et al.
(2003) and our late-time datum is consistent with 90◦, a
quantity predicted in the more commonly accepted mod-
els for the polarizations of GRB jets (GL99; Sari 1999).
However, in contrast with the prediction, the transition
between the two angles is not sharp, but rather smooth.
This suggest the presence of an additional polarizing com-
ponent, as we will discuss below.
3.2. Spectropolarimetry
We have shown in Fig. 2 the result of our reduction of
the spectropolarimetric observation publicly available at
ESO. Spectropolarimetric observations of OTs are impor-
tant since they allow us, in principle, to single out which
component is contributing more to the observed polariza-
tion: the OT itself, the host ISM, or the MW ISM. This is
possible since the three components have different wave-
length dependencies.
The OT polarization is supposed to be wavelength in-
dependent, at least in the limited spectral range investi-
gated here. ISM polarization is instead wavelength depen-
dent. It follows a "Serkowski law":
P (λ) = Pmax exp(cid:20)−K ln2(cid:18) λmax
λ (cid:19)(cid:21) ,
(1)
where Pmax is the maximum induced polarization, ob-
tained at λmax. Experimentally, 0.34 µm <∼ λmax <∼
0.9 µm, and this parameter is considered to be a mea-
sure of the size of the polarizing grains: the larger λmax,
Table 2. Results of the simple modelling of the spectropo-
larimetric result. † The constant polarization value is given
for the OT model. ‡ The value λmax = 0.34 µm has been
set as a lower limit to the parameter (see text).
Pmax(%)
1.88 ± 0.05
Model
OT†
MW ISM 1.95 ± 0.1
Host ISM 2.1 ± 0.15
ϑ (◦)
118 ± 1 --
118 ± 1
118 ± 1
λmax (µm)
0.52 ± 0.05
0.34‡
χ2/d.o.f.
45.8/44
50.5/43
57.2/43
the larger the grains. For the numerical value of the coef-
ficient K, we follow the more recent study of Martin et al.
(1999):
K = (cid:26) 1.66 λmax
λ ≥ λmax;
−0.59 + 2.56 λmax λ < λmax.
(2)
To model the polarization results shown in Fig. 2, we
have assumed first that the entire polarization is due to
only one of the three possible components (i.e. OT, MW,
host), taking into account the wavelength dependence of
the ISM polarization as detailed above, and calculating the
effect of the host-ISM in the host rest frame5. As reported
in Tab. 2, the quality of the data allows us to exclude a
dominant role of the host ISM in producing the observed
polarization, but it is not possible to identify a dominant
component between MW-ISM and intrinsic. The first two
fits are both acceptable (χ2/d.o.f. < 1.2), with a slight
preference for the OT model, while the host-ISM fit yields
a significantly worst χ2. Performing the fit only on the
high quality dataset (filled symbols in Fig. 2) yields sta-
tistically indistinguishable results. However, since the field
stars do show a moderate degree of polarization, we know
that the observed polarization cannot be entirely due to
the OT itself. Furthermore, at all frequencies the polar-
ization position angle is very similar to that of the field
stars in the imaging polarimetry (see above).
5 The application of the Serkowski law to high-redshift sys-
tems is not directly supported by observations, given the in-
trinsic difficulty in performing spectropolarimetric studies of
distant objects. It is however at least partly justified by the
observation of interstellar polarization following a Serkowski
curve in SN 1986G in Centaurus A, at z = 0.00183 (Hough et
al. 1987).
6
Lazzati et al.: Polarization of GRB 021004
Fig. 4. Effect of a polarizing ISM on the theoretical polarization of afterglows. The theoretical curves (top: polarization
level; bottom: position angle) from GL99 for ϑo/ϑj = 0.5 are shown, in both panels, as solid lines (pISM = 0). Lines with
different styles show the effect of the selective extinction for ISM with small, intermediate and comparable polarization
with respect to the OT maximum polarization. The position angles of the ISM are reported relative to the initial angle
of the OT. The left panel shows the effect of an ISM that induces polarization with a position angle ϑISM = −19◦,
while the right panel shows an ISM with ϑISM = 60◦.
3.3. Combined modelling
Since the polarization levels are small and different effects
seem to contribute at comparable degree, it is mandatory
to perform a combined modelling of the imaging and spec-
tral polarimetry, combining the effects of the OT polariza-
tion and of the MW ISM selective extinction. The need of
at least two polarizing components is due to the presence
a) of time variability of the measured polarization (OT
component) and b) of significant polarization in the field
stars (MW ISM component).
3.3.1. Transmission of variably polarized light in a
polarizing ISM
In order to combine the effects of ISM selective extinc-
tion with the evolving intrinsic polarization of the OT,
we adopt a Mueller calculus approach (see e.g. di Serego
Alighieri 1997 and references therein). In this formalism,
the transmitted Stokes vector S′ ≡ (I ′, Q′, U ′, V ′) is com-
puted from the incident one S ≡ (I, Q, U, V ) through a
matrix, called "Mueller matrix", which incorporates all
the properties of the transmitting medium: S′ = M · S.
Since developing a complete treatment of the polarizing
properties of the ISM is far beyond the scope of this paper,
here and in the following we adopt a simplified version of
the Mueller matrix which is correct for a non-birefringent
dichroic medium. With this simplification we assume that
the ISM is not able to induce circular polarization to any
incident light, either polarized or not. In fact, the ISM
does induce a small degree of circular polarization even in
unpolarized sources (Martin & Angel 1976), which indi-
cates that a non-coaxial birefringent and dichroic medium
should be considered. However, the optical properties of
such a medium are described by five parameters (extinc-
tion, orientation of the optical axes of birefringence and of
dichroism and the two respective refraction indexes) while
we can constrain only four parameters observationally (ex-
tinction, induced linear polarization, position angle and
induced circular polarization). It is therefore not feasible
to derive a complete Mueller matrix without going into a
detailed modelling of the structure and geometry of dust
grain and their alignment.
Consider a dichroic ISM that induces a polarization
pISM ≡ (q2 + u2)1/2 on unpolarized stars. Its Mueller ma-
trix has then the form:
M = e−τ
q
p2
1
q q2+Au2
u qu(1−A)
0
p2
0
ISM
ISM
u
qu(1−A)
p2
ISM
u2+Aq2
ISM
p2
0
0
0
0
A
(3)
where A ≡ p1 − p2
ISM, and e−τ is the opacity of the
medium to non-polarized radiation. In order to conserve
energy (i.e. not to have an increased transmitted inten-
sity), this parameter must satisfy:
e−τ ≤
1
1 + q + u
,
(4)
the equality holding for a perfect polarizing medium, i.e.
one that does not absorb any radiation completely po-
Lazzati et al.: Polarization of GRB 021004
7
larized with its same angle. It can also be easily shown
that if pISM ≪ 1 and Q, U, V ≪ 1 the matrix com-
putation is equivalent to a simple sum of the incident
and ISM normalized Stokes parameters: Q′ ≈ q + Q,
U ′ ≈ u + U . Formally, the Mueller matrix in Eq. 3 is
not valid in the case of a birefringent medium. It can
however be shown that it is a good approximation (de-
viations < 10%) if the incoming light is moderately po-
larized (P < 10%), the linear and circular interstellar po-
larizations are small (pISM < 10% and vISM/pISM < 10−2;
Martin & Angel 1976) and the circular polarization in-
duced on polarized sources is small (V /P < 10−1 for ac-
tive galactic nuclei; Landstreet & Angel 1972).
In order to exemplify the effect of a polarizing ISM on
the intrinsic OT polarization, we adopt the polarization
model by GL99. This model predicts a polarization curve
characterised by two distinct peaks, whose absolute inten-
sity depend on the angle between the line of sight and the
axis of the fireball (assumed to be collimated in a jet).
The polarization position angle is rotated by 90 degrees
at the time of null polarization between the two peaks.
This model is illustrated by the solid lines in Fig. 4 (i.e.
no ISM polarization). With different line-styles, this figure
also shows how the predicted polarization degree and po-
larization angle change in time once modified by some in-
tervening polarizing ISM. All angles are reported relative
to the initial OT polarization angle. In the left panel the
case relative to this burst is shown (ϑ = −18◦; see below),
while in the right panel a larger misalignment between the
OT and ISM angles is shown. The figure shows that the
presence of the ISM modifies quite substantially the ob-
served polarization, especially at times when POT ≈ PISM.
Note also that the presence of the polarizing ISM makes
the position angle to vary smoothly, instead of sharply.
Fig. 5. Time resolved modelling of the polarization and
position angle of GRB 021004. The first three and the fifth
datum are taken from Rol et al. (2003). Our NIR upper
limit is not included. The solid line show the best fit ob-
tained from the whole dataset, with a free break time and
free properties of the ISM polarization. Even with this
extra freedom a good fit cannot be obtained due to the
rapidity of the evolution of the polarization (see text for
more details). The dashed line shows instead the best fit
model obtained by modelling only the last four data. In
this case the model yields acceptable χ2/d.o.f. = 6.3/4
and a best-fit break time comparable to the one derived
from the lightcurve fitting (see text for more details). The
vertical dotted line shows the break time tb = 4.74 d de-
rived from lightcurve modelling (Holland et al. 2003).
3.3.2. Modelling of the polarization curve
To model the time-resolved polarization measurements in
the framework of available models (GL99) with a polariz-
ing ISM we perform a fit to the polarization curve propa-
gating the intrinsic OT polarization through the MW ISM
with the use of the Mueller matrix derived in Eq. 3. The
model has 6 degrees of freedom: the initial position an-
gle of the OT intrinsic polarization, the off-axis angle of
the line of sight, the degree of alignment of the magnetic
field, the jet break time and the ISM q and u parameters.
The last three parameters can be constrained with the ob-
servations, but we let them free to vary here to check a
posteriori the agreement of the derived values with the ob-
servations. In addition, we adopt models with and without
sideways expansion of the jet (GL99; Sari 1999; Rossi et al.
2003, in prep). A particularly important issue is to check
whether the prediction of the models are consistent with
the position angle rotation detected by Rol et al. (2003).
The fit is performed on the q and u parameters rather
than on p and ϑ due to their better statistical properties.
The result of the fit is that it is not possible to model
the ∼ 90◦ rotation of the position angle in the framework
of the proposed models, not even with the addition of a
polarizing ISM with free properties. In fact, the smallest
χ2 that can be obtained is χ2 = 43 for 8 degrees of free-
dom (for a non sideways expanding jet). In addition, the
fit formally yields a best break time tj = 0.25 d, in dis-
agreement with the value tj = 4.74+0.14
−0.8 inferred from the
lightcurve modelling (Holland et al. 2003). Also, the best
fit excludes the presence of a sizably polarizing ISM, in
disagreement with the observation of polarization of the
stars in the field of GRB 021004. This could be explained
with a polarizing ISM in the host galaxy with properties
opposite to those of our ISM. Such a possibility would
however require an unacceptable fine tuning and is ruled
out by the modelling of the spectropolarimetric observa-
tion (see § 3.2). The formal best fit model, re-converted in
polarization and position angle, is shown in Fig. 5 with a
solid line.
There is one important issue, however, that makes this
result non conclusive. The models for polarization that we
have used in the fit above are derived under the assump-
8
Lazzati et al.: Polarization of GRB 021004
tion that the fireball and the ISM are homogeneous. In
this case, the total lightcurve should be characterised by
a smooth broken power-law decay. This is not the case for
GRB 021004. Its lightcurve, as shown in Fig. 1 (see also
Lazzati et al. 2002b), shows prominent bumps overlaid on
this power-law. There are several possible explanations for
these bumps: inhomogeneities in the ISM (Lazzati et al.
2002b), inhomogeneities in the fireball (Nakar et al. 2003)
or delayed injection of energy from the central engine (Fox
et al. 2003). While in the latter hypothesis the polariza-
tion curve should not be affected, inhomogeneities, either
in the fireball or in the ISM, can produce a polarization
signal by breaking the symmetry of the fireball emission
at early times (Granot & Konigl 2003). Such polarization
would have a randomly oriented position angle, depend-
ing on the azimuthal location of the inhomogeneity with
respect to the line of sight. A similar effect is predicted in
the case of micro-lensing events (Loeb & Perna 1998). If
we look more closely at Fig. 1, we note that the first three
measurement of Rol et al. (2003) lie on top of a major
rebrightening, while the remaining four observations were
performed in relatively unaffected time intervals. Only the
last four measurements should therefore closely follow the
theoretical model.
To check this idea, we have performed the same mod-
elling described above only on the final four points, fixing
the ISM polarizing properties to that of the stars in the
field of GRB 021004 in order to limit the number of free
parameters. Indeed, the last four points can be successfully
described by the model, with a break time tj = 3 ± 1 days,
in good agreement with the tj = 4.74+0.14
−0.8 d obtained by
Holland et al. (2003) from the break in the lightcurve. The
best fit model, which has χ2 = 6.3 for 4 degrees of free-
dom, is shown with a dashed line in Fig. 5. The fit requires
a large degree of alignment of the magnetic field of 75%
and a moderate off-axis line of sight ϑo/ϑj = 0.45. The
best fit model is obtained for a non sideways expanding
jet. A sideways expanding jet cannot however be rejected,
with χ2/d.o.f. = 7.2/4. This ambiguity is due to the fact
that, according to the best fit model, the measured points
lie in the first peak of the polarization curve, where the
sideways expansion has only a marginal effect. The best fit
model requires a misalignment of (−19 ± 1)◦ between the
OT initial position angle and the ISM polarization. The
effect of such a misalignment is shown in the left panel of
Fig. 4.
3.4. Complications
Additional complications can be envisaged in locally red-
dened afterglows. If in fact a strong polarization is induced
by dust in the close vicinity of the burst explosion site,
two effects can take place. First, dust destruction by the
burst prompt and afterglow emission (Waxman & Draine
2000; Perna & Lazzati 2002) can imprint a strong tempo-
ral evolution in the host ISM polarization. This is however
unlikely to be relevant for our observations, since the typ-
ical time scale for this process is of the order of minutes,
rather than days. Second, a patchy absorber may obscure
in a different way different portions of the fireball (e.g.
Wang et al. 2003), altering one of the key assumptions of
the models. None of these effects should alter dramatically
the lightcurve, but a random noise could be overlaid on the
smooth theoretical evolution of the polarization predicted
by the models.
In addition, a preferential direction for the polariza-
tion, different from the one defined by the plane contain-
ing the jet axis and the line of sight, can be defined by
the presence of an interstellar magnetic field of sufficient
magnitude. Such a possibility has been recently studied by
Granot & Konigl (2003). Qualitatively, the effect of such
a pre-existing field is not different from that of a polar-
izing ISM. Due to the presence of a second asymmetry,
the evolution of the polarization angle is smooth rather
than sharp, and the external component dominates the
observed polarization properties when the intrinsic OT
one is small. Such additional degrees of freedom were not
necessary in our last fit (nor they can explain the early
∼ 90◦ angle rotation). We stress however that, since all
our data were observed before the jet break, we cannot
exclude the presence of an ISM magnetic field with a well
defined orientation.
4. Summary and conclusions
We have presented multi-time and multi-filter observa-
tions of the polarization of the afterglow of GRB 021004
performed with ESO -- VLT and TNG as well as our anal-
ysis of the publicly available ESO -- VLT spectropolarimet-
ric observation. The interpretation of the observations is
complex since none of the polarizing mechanisms that con-
tribute to the observed polarization seems to clearly dom-
inate over the others. We therefore adopt this afterglow as
a case study to investigate and describe these three main
effects: intrinsic (and time varying) OT polarization, host
ISM polarization, and MW ISM polarization. By mod-
elling the spectropolarimetric and time-resolved imaging
polarimetry we were able to get rid of the contribution of
the host ISM, while OT and MW ISM polarizations seem
to play an intertwined role, one dominating over the other
at different times.
To perform a detailed time dependent modelling of
the polarization and position angle evolutions, we imple-
mented our data-set with the four observations of Rol et
al. (2003). The complete dataset is particularly interesting
since a sizable rotation of the position angle is present. Our
attempt to fit the polarization data within the framework
of GL99 models was however a failure, given the short
timescale of the evolution and the lack of an appreciable
break in the lightcurve at the time of the position angle
evolution. Since the angle rotation is associated to one
of the rebrightening events in the lightcurve, this burst
is not suited for a comparison with models, which are
computed for homogeneous fireballs producing featureless
lightcurves. A possible explanation within exsisting mod-
Lazzati et al.: Polarization of GRB 021004
9
Granot J., Panaitescu A., Kumar P., & Woosley S.E. 2002,
ApJ, 570, 61
Granot J., & Konigl A., 2003, ApJ, 594, L83
Henden A. 2002a, GCN 1583
Henden A. 2002b, GCN 1630
Hjorth J., Bjornsson G., Andersen M.I., et al. 1999, Science,
283, 2073
Holland S.T., Weidinger M., Fynbo J.P.U., et al. 2003, AJ,
125, 2291
Hough J.H., Bailey J.A., Rouse M.F., Whittet D.C.B. 1987,
MNRAS, 227, 1
Klose S., Stecklum B., & Fischer O. 2001, in "Gamma-Ray
Burst in the Afterglow Era", eds. E. Costa, F. Frontera,
J. Hjorth. Berlin Heidelberg: Springer, 2001, 188
Lazzati D., Covino S., Ghisellini G. 2002a, MNRAS, 330, 583
Lazzati D., Rossi E., Covino S., Ghisellini G., & Malesani D.
2002b, A&A, 396, L5
Landstreet J.D., & Angel J.R.P. 1972, ApJ, 174, L127
Loeb A., & Perna R. 1998, ApJ, 495, 597
Magalhaes A.M., Pereyra A., Dominici T., & Abraham Z. 2003,
GCN 2163
Martin P.G., & Angel J.R.P. 1976, ApJ, 207, 126
Martin P.G., Clayton G.C., & Wolff M.J. 1999, ApJ, 510, 905
Masetti N., Palazzi E., Pian E., et al. 2003, A&A, 404, 465
Matheson T., Garnavich P.M., Foltz C., et al. 2002, ApJ, 582,
L5
Mirabal N., Halpern J.P., Chornock R., & Filippenko A.V.
2002, GCN 1618
Møller P., Fynbo J.P.U., Hjorth J., et al. 2002, A&A, 396, L21
Nakar E., Piran T., & Granot J., 2003, New Ast., 8, 495
Oliva E. 1997, A&AS, 123, 589
Pandey S.B., Sahu D.K., Resmi L., et al. 2002, BASI submit-
ted; astro-ph/0211108
Perna R., & Lazzati D. 2002, ApJ, 580, 261
Rol E., Wijers R.A.M.J., Vreeswijk P.M., et al. 2000, ApJ, 544,
707
Rol E., Wijers R.A.M.J., Fynbo J.P.U., et al. 2003, A&A, 405,
L23
Sari R. 1999, ApJ, 524, L43
Schaefer B.E., Gerardy C.L., Hoflich P., et al. 2003, ApJ, 588,
387
Schlegel D.J., Finkbeiner D.P., & Davis M. 1998, ApJ, 500,
525
Serkowski K., Mathewson D.L., & Ford V.L. 1975, ApJ, 196,
261
Shirasaki Y., Graziani C., Matsuoka M., et al. 2002, GCN 1565
Uemura M., Kato T., Ishioka R., & Yamaoka H. 2003, PASJ,
55, L31
Wang L., Baade D., Hoeflich P., & Wheeler J.C. 2003, ApJL
submitted (astro-ph/0301266)
Wardle J.F.C., & Kronberg P.P. 1974, ApJ, 194, 249
Waxman E., & Draine B.T. 2000, ApJ, 537, 796
Wijers R.A.M.J., Vreeswijk P.M., Galama T.J., et al. 1999,
ApJ, 523, L33
els is that the early time polarization is dominated by a
local enhancement of the fireball emission, which breaks
the symmetry of the fireball producing a polarized signal.
Indeed, the late time polarization data can be successfully
and consistently described by the models. We therefore
suggest that the lightcurve bumps are due to local events
on the fireball surface, such as inhomogeneities in the ISM
(Lazzati et al. 2002b) or within the fireball itself (Nakar
et al. 2003). Refreshed shocks (Fox et al. 2003) are in-
stead unable to account for the combined lightcurve and
polarization evolution and can be rejected.
The quality of the data, and in particular the lack of
late time measurements) does not allow us to pin down
the geometry and dynamic of the outflow, but we have
shown that, with good quality spectropolarimetry and
multi-time/multi-filter imaging polarimetry, it is in prin-
ciple possible to disentangle the three effects and get a
hold on the intrinsic polarization and on the structure
and dynamics of GRB outflows. The added value of such
a measurement would be the study of the polarizing prop-
erties of dust in high redshift galaxies, a poorly studied
property of such an important component of high redshift
objects.
Acknowledgements. We thank the TNG and Paranal ESO
staffs for the professional, kind and reliable support. We thank
Evert Rol for providing its results in table form. This work
was partly made using public data from the ESO VLT Science
archive. The work of Scott Barthelmy in maintaining the
GCN system is invaluable and greatly appreciated. DL ac-
knowledges support from the PPARC postdoctoral fellowship
PPA/P/S/2001/00268.
References
Barth A.J., Sari R., Cohen M.H., et al. 2003, ApJ, 584, L47
Bersier D., McLeod B., Garnavich P., et al. 2003a, ApJ, 583,
L63
Bersier D., Stanek K.Z., Winn J., et al. 2003b, ApJ, 584, L43
Bjornsson G., Hjorth J., Pedersen K., & Fynbo J.P.U. 2002,
ApJ, 579, L59
Covino S., Lazzati D., Ghisellini G., et al. 1999, A&A, 348, L1
Covino S., Malesani D., Ghisellini G., et al. 2002a, GCN 1498
Covino S., Lazzati D., Malesani D., et al. 2002b, A&A, 392,
865
Covino S., Malesani D., Ghisellini G., et al. 2003a, A&A, 400,
L9
Covino S., Ghisellini G., Malesani D., et al. 2003b, GCN 2167
Covino S., Ghisellini G., Lazzati D., & Malesani D. 2003c, in
"Gamma-Ray Bursts in the Afterglow Era", Roma (Sept.
16 -- 20), in press (astro-ph/0301608)
Covino S., Ghisellini G., Malesani D., et al. 2003d, GCN 1909
Devillard N. 1997, The Messenger, 87, 19
di Serego Alighieri S. 1997,
in "Instrumentation for Large
Telescopes", eds. J.M. Rodriguez Espinosa, A. Herrero,
F. Sanchez, Cambridge University Press, 287
Efimov Y., Antoniuk K., Rumyantsev V., & Pozanenko A.
2003, GCN 2144
Fox D.W. 2002, GCN 1564
Fox D.W., Yost S., Kulkarni S.R., et al. 2003, Nature, 422, 284
Ghisellini G., & Lazzati D. 1999, MNRAS, 309, L7 (GL99)
|
astro-ph/0304479 | 1 | 0304 | 2003-04-26T18:13:15 | XMM-Newton observations of Nova LMC 2000 | [
"astro-ph"
] | We report on three X-ray observations of Nova LMC 2000 with XMM at 17, 51 and 294 days after the maximum, respectively. X-ray spectral fits show a concordant decrease of the absorbing column and the X-ray luminosity. No supersoft X-ray emission is detected. The mass of the ejected shell is determined to be (less than) 7.5*E-5 msun. Though data are sparse, one interesting correlation becomes visible: sources with a long-duration supersoft X-ray phase have shorter orbital periods than those with short or no supersoft X-ray phase. This can be understood considering that (i) enough matter has to be accreted in order to ignite the H burning, and (ii) that H burning ceases when the mass of the remaining material (after shell ejection and burning) drops below a certain limit under which the temperature at the bottom of the envelope is too low for the shell burning to compensate the energy loss from the surface. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. nlmc00xmmed
(DOI: will be inserted by hand later)
November 16, 2018
3
0
0
2
r
p
A
6
2
1
v
9
7
4
4
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
XMM-Newton observations of Nova LMC 2000⋆
J. Greiner1, M. Orio2
,
3, and N. Schartel4
1 Max-Planck-Institut fur extraterrestrische Physik, 85741 Garching, Germany
2 INAF - National Institute for Astrophysics, Osservatorio Astronomico di Torino, Strada Osservatorio 20, 10025
Pino Torinese, Italy
3 Department of Astronomy, 475 N. Charter Str., Madison WI 53706, USA
4 XMM-Newton Science Operations Centre, ESA, Villafranca del Castillo, P.O. Box 50727, 28080 Madrid, Spain
Received 27 February 2003 / Accepted 23 April 2003
Abstract. We report on three X-ray observations of Nova LMC 2000 with XMM-Newton at 17, 51 and 294 days
after the maximum, respectively. X-ray spectral fits show a concordant decrease of the absorbing column and
the X-ray luminosity. No supersoft X-ray emission is detected. The mass of the ejected shell is determined to
be (less than) 7.5 × 10−5 M⊙. Though data are sparse, one interesting correlation becomes visible: sources with
a long-duration supersoft X-ray phase have shorter orbital periods than those with short or no supersoft X-ray
phase. This can be understood considering that (i) enough matter has to be accreted in order to ignite the H
burning, and (ii) that H burning ceases when the mass of the remaining material (after shell ejection and burning)
drops below a certain limit under which the temperature at the bottom of the envelope is too low for the shell
burning to compensate the energy loss from the surface.
Key words. Stars: individual: N LMC 2000 -- Stars: mass loss -- novae, cataclysmic variables -- X rays: stars --
binaries: close
1. Introduction
Nova LMC 2000 was discovered in mid-July 2000 (Liller
2000). Early observations between June 29.68 and July
1.68 have detected the nova, but no brightness estimates
are available (Bond & Kilmartin 2000), since all images
except the first one on June 29.68 show the nova to be
saturated (Bond, priv. comm.). This leaves the exact time
and magnitude of the maximum unknown except for the
constraint that it was still rising on June 29.68 (Bond,
priv. comm.). But even if the observed mV = 11.2 mag
on 12 July 2000 was the maximum, with a corresponding
to MV = −7.5 mag, Nova LMC 2000 was among the 20%
brightest novae.
Optical spectroscopy on 15 July 2000 shows emission
lines of the Balmer series, several Fe multiplets and Na I D,
suggesting a "Fe II" nova about one week after maximum
(Duerbeck & Pompei 2000). If this time estimate is cor-
rect, the maximum of the nova could have been mV = 10.5
mag (backwards extrapolation to 8 July 2000) correspond-
ing to MV = −8.2 mag.
Spectroscopy of this ONeMg nova (Shore 2002) also
revealed weak absorption components in the Balmer lines
at 1900 km/s on July 15 (Duerbeck & Pompei 2000), and
Send offprint requests to: J. Greiner, [email protected]
⋆ Partly based on observations collected at the European
Southern Observatory, Chile under proposal 66.D-0391
a FWZI of the Balmer lines on July 14 and 15 of 2200±250
km/s (Hearnshaw & Yan Tse 2000). Spectra obtained with
HST on August 19/20, 2000, strongly resemble those of
the fast nova V382 Vel (t3 = 9 days; Della Valle et al. 2002)
at 2 months after visual maximum (Shore et al. 2000).
Only 4 out of about 100 classical novae observed with
ROSAT have been found to exhibit a supersoft phase (see
Orio & Greiner 1999, Orio et al. 2002). However, theoret-
ically one expects that each nova should pass through a
phase of soft X-ray emission during the later stages of
decline (e.g. MacDonald et al. 1985): During the post-
maximum stage, at constant bolometric luminosity, the
photosphere progressively retreats as the residual hydro-
gen envelope is depleted, and the effective temperature
rises up to the soft X-ray region. It is pretty much unclear
yet what determines the appearance of supersoft X-ray
emission. It has been argued (Truran 2002) that the on-
set of the supersoft phase is the same for all novae, about
6-8 months, and that the duration of the supersoft X-
ray phase is the burning timescale of the white dwarf, i.e.
proportional to the mass of the white dwarf (Truran &
Glasner 1995, Vanlandingham et al. 2001). While the first
two supersoft novae, GQ Mus ( Ogelman et al. 1993) and
V1974 Cyg (Krautter et al. 1996), are consistent with this
suggestion, observations of other recent novae imply that
the picture is more complicated.
2
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
Table 1. Observation log
Observation Interval
(UT)
Time after
Max. (days)
Expo.-Time Rev. Count rate
(cts/ksec)
(ksec)
HR1
HR2
HR3
2000-07-25 22:48 -- 2000-07-26 03:19
2000-08-28 13:53 -- 2000-08-28 16:40
2001-03-29 19:10 -- 2001-03-29 21:22
17
51
294
16.26
10.00
10.54
115
132
239
5.0±0.6
21.0±1.6
<1.0
--
0.70±0.07
1.00±2.00
-- 0.56±0.06
0.00±0.11
-- 0.66±0.16
--
--
--
(1) The hardness ratios are defined as HR1 = (B − A)/(B + A), HR2 = (C − B)/(B + C) and HR3 = (D − C)/(D + C), where
A(0.2 − 0.5 keV), B(0.5 − 2.0 keV), C(2.0 − 4.5 keV), D(4.5 − 7.5 keV), are the counts in the given energy range.
Here we report the results of a sequence of XMM-
Newton observations of N LMC 2000 aimed at finding and
characterizing the supersoft X-ray component. We also re-
port a contemporaneous optical observation.
2. Observations and Results
2.1. XMM-Newton observations
Observations of Nova LMC 2000 were performed at three
occasions (see Tab. 1), about 17, 51 and 294 days after
the likely maximum on 8 July 2000.
Throughout all observations the thin blocking filter
was used. In the following, we primarily deal with EPIC-
pn (Struder et al.2001) part of the X-ray data, and the
optical monitor (OM; Mason et al. 2001) data.
2.1.1. XMM-Newton EPIC-pn data
After a first inspection of the X-ray data, a source de-
tection run was done in the 0.5 -- 7 keV band, as well as
in narrow bands to derive hardness ratios. These hard-
ness ratios, the count rates in the full band and the 3σ
upper limit for the non-detection during the March 2001
observation are given in Tab. 1 and show the strong inten-
sity and spectral variability of the X-ray emission of Nova
LMC 2000.
For the spectral analysis, single and double events
away from the edges of the CCD or bad pixels were ex-
tracted in a source region of 19′′, and a background region
of outer diameter of 30′′.
Because of the larger number of collected photons we
first fitted the X-ray spectrum obtained on 28 August
2000. Before the extraction of source and background pho-
tons, a time interval of ∼2 ksec was excluded because of
high background. Since the spectrum lacks any soft com-
ponent, we applied a bremsstrahlung and a Raymond-
Smith model to the data. Both models give satisfactory
fits, and the fit parameters are given in Tab. 2. During
the observation on 28 August 2000, the unabsorbed X-ray
luminosity was 2×1034 erg/s (0.01 -- 20 keV; assuming a dis-
tance of 55 kpc). Note that despite the (best-fit) absorbing
column of (2 -- 3)×1021 cm−2, a supersoft component with
a characteristic temperature of 30 -- 50 eV and a luminosity
of 1036 -- 1038 erg/s would have been easily detectable.
of counts, the temperature in both models is not con-
strained. Thus, in a second iteration, we fixed the tem-
perature to the value as derived from the August 28, 2000
observation. The best-fit parameters of both fits are also
given in Tab. 2. The unabsorbed bolometric luminosity of
Nova LMC 2000 during the July 25/26, 2000 observation
was (5.0±0.3)×1034 (D/55 kpc)2 erg/s, where the error
includes also the difference between the two models. In a
third iteration we tested a fit with the absorbing column
fixed to the value of the July 2000 observation in order to
check whether a temperature change could mimic the vari-
ation in absorbing column. However, this does not provide
an acceptable fit (reduced χ2 of 2.5).
For the X-ray non-detection on 29 March 2001, assum-
ing the same spectral models and no intrinsic absorption
through the ejected nova shell we derive a 3σ upper limit
for the luminosity of L < 1.1 × 1033 (D/55 kpc)2 erg/s for
the case of only galactic foreground absorption (7×1020
cm−2) or L < 1.4 × 1033 (D/55 kpc)2 erg/s for the case of
galactic foreground plus total LMC absorption (15×1020
cm−2; Luks 1994).
2.1.2. XMM-Newton Optical Monitor data
The OM was used with a variety of filters, and also the
grism was used in two occasions. In particular, we ob-
tained (i) two 5 ksec grism 2 (visual spectrum) exposures,
a sequence of 5 exposures with the UVW1 filter, and a
sequence of 4 exposures with the UVW2 filter, each with
940 sec) in the first XMM-Newton observation of Nova
LMC 2000 (revolution 115); (ii) a sequence of 5 exposures
with the V band filter with 1 ksec each, followed by a 5
ksec grism 2 exposure (second XMM-Newton observation;
rev. 132); (iii) a sequence of 5 exposures of 1 ksec with the
U band filter in the last XMM-Newton observation (rev.
239).
Unfortunately, the grism data are hardly usable since
the window of the CCD was set too small. The photometry
in the different filters is summarized in Tab. 3. The central
wavelengths for the UV filters are 2910 A for the UVW1
and 2120 A for the UVW2, respectively.
2.2. Optical photometry
We then used the same models to also fit the July
25/26 observation. However, due to the smaller number
On 4 December 2000 we obtained three V band expo-
sures with 90 sec exposure time each using DFOSC at
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
3
Fig. 1. X-ray spectrum of Nova LMC 2000 as observed with XMM-Newton on 28 August 2000, modelled with a
bremsstrahlung (left) and Raymond-Smith model (right). The top panel shows the count rate spectrum and the
residuals, while the lower panels show the unfolded photon spectrum.
Table 2. Spectral fit results of the July and August 2000 observations
Date
NH
(1022 cm−2)
kT
(keV)
Normalization
(ph keV−1 cm−2s−1)
Flux (bolometric)
(erg cm−2s−1)
χ2 / dof
25/26 Jul 2000
25/26 Jul 2000
28 Aug 2000
3.23±1.23
7.00±2.57
0.29±0.19
199±100
1.6 fixed
1.6±1.2
1.1E-05
3.6E-05
1.6E-05
bremsstrahlung model
25/26 Jul 2000
25/26 Jul 2000
28 Aug 2000
3.40±1.10
7.65±2.33
0.20±0.15
64±50.0
2.1 fixed
2.1±1.0
2.4E-05
7.5E-05
2.3E-05
Raymond-Smith model
7.5E-14
13.0E-14
5.5E-14
6.9E-14
14.4E-14
4.5E-14
0.78
1.28
0.94
0.79
1.00
0.92
the 1.5 m Danish telescope at La Silla (ESO, Chile). The
MAT/EEV 44-82 CCD chip with 15 µm pixels has a plate
scale of 0.′′4/pixel. The seeing was 1′′. The photometric
standard RU 149, observed at nearly the same airmass
as N LMC 2000, was used for the absolute flux calibra-
tion. Due to the brightness of the nova and the relatively
crowding-free surrounding (see Fig. 2), aperture photom-
etry was done within the MIDAS package after standard
flatfield and bias correction. We measure V = 16.20±0.03
mag independently in each of the three images for Nova
LMC 2000. After adding this to the early measurements
by VSNET observers (Fig. 3), a comparison with light
curve models of Hachisu & Kato suggests that Nova LMC
2000 was possibly still in the plateau phase during our
optical observation on 4 December 2000.
A crude estimate of the decay rate based on the
VSNET data gives t2 ∼ 9.0 days and t3 ∼ 22 days, with
an error of ±2 days depending on the actual occurrence
of the not observed optical maximum (where t2 and t3 are
4
5
5
3
1
Fr, 21 Jun 2002 13:45:06
MIDAS version: 99NOV
Tu, 18 Feb 2003 15:02:12
MIDAS version: 02FEB
8
8
1
2
.
0
7
-
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
0
0
1
1
802
5
2
5
2
.
0
7
-
1057
81.266
81.2322
Fig. 2. Finding chart of N LMC 2000 in the V band (left; marked with two dashes near the image center) and at 2120
a (right). The V band image was taken on 2000 Dec. 4 with DFOSC at the 1.5 m Danish telescope at La Silla (ESO).
We measure RA = 5h25m01.s1 and Decl. = -70◦14′17′′ (equinox 2000.0) with an error radius of 2′′. The field size is
1.′7×1.′7. The right image is the sum of 4 images with 1000 sec exposure each, taken by the OM onboard XMM in the
UVW2 filter on 25/26 July 2000, where the nova is the by far brightest object in the image. The field size is 2.′3×2.′3.
North is up and East to the left.
Table 3. XMM/OM photometry
3.2. The mass of the ejected envelope
Date
filter
countrate brightness(1)
cts/sec
mag
flux(1)
erg/cm2/s/A
2000-07-25 UVW1
UVW2
2000-08-28
2001-03-29
V
U
201.4
16.0
10.6
9.5
11.41±0.01
11.62±0.01
15.24±0.02
15.70±0.02
1.04E-13
1.24E-13
3.05E-15
2.24E-15
(1) Brightness and flux are given without correction for galac-
tic, LMC and nova-intrinsic extinction, since the latter two
values have uncertainties much larger than the measurement
error.
the times during which the brightness decays by two and
three magnitudes, respectively).
3. Discussion
3.1. Luminosity and absorption change
The 0.1 -- 10 keV X-ray emission observed from Nova LMC
2000 at 17 and 51 days after the explosion is well de-
scribed by a bremsstrahlung or Raymond-Smith model.
Spectral fitting reveals that during the first observation
the absorbing column was much larger, indicating intrin-
sic absorption due to the ejected shell. This is independent
of the model used in the fitting. Despite the smaller count
rate in the first observation with respect to the second,
the derived luminosity was largest in the first observation.
The large excess absorption determined from the first ob-
servation can be used to estimate the mass of the ejected
shell, assuming no clumpiness (and a filling factor of 1
though this is likely an overestimate), i.e. that the mea-
sured column density is representative for all spatial di-
rections around the nova system. At the time of the
first observation, the expansion velocity was ∼2000 km/s
(Duerbeck & Pompei 2000, Hearnshaw & Yan Tse 2000),
and thus the shell radius was R ∼ 3 × 109 km, if no sub-
stantial deceleration has occurred during the first 17 days
after the nova explosion. Using a mean column density of
neutral hydrogen of 5 × 1022 cm−2 (see Tab. 2), a mass
of the shell of ∆Mej = 7.5 × 10−5 M⊙ is obtained. As a
kind of consistency check one can also consider the later
decrease of the absorbing column between the first and
second XMM-Newton observation. Since the time differ-
ence, and thus the radius difference, is a factor of 3, the
absorbing column should have reduced by a factor of 9
if the shell had expanded with constant velocity. This is,
within the errors, consistent with the measured change in
column density (see Tab. 2).
An independent estimate of the ejected shell mass can
be derived from the measured decay time t2 and the rela-
tion log ∆Mej = 0.274 × log t2 − 4.355 (Della Valle et al.
2002; note that their formula as well as their figure miss a
factor of 10−5 in ∆Mej). This yields ∆Mej = 8×10−5 M⊙,
in good agreement with the above estimate from the mea-
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
5
In an attempt to understand the reason for the largely
varying duration of the supersoft X-ray phase in differ-
ent novae we have compiled some information on novae
with observed supersoft X-ray phase or very constraining
observational limits (Tab. 4).
Fig. 3. Light curve of Nova LMC 2000 as measured by var-
ious observers and provided to VSNET (filled circles). Our
ground-based observation from Dec. 2000 is marked by a
filled triangle, and the XMM/OM observation in August
2000 by a filled lozenge. The cross marks the possible max-
imum as determined by a backwards extrapolation (see
Introduction). Vertical dashes at the top mark the times
of the XMM-Newton observations. The dashed line visu-
alizes the decay curve of the recurrent nova CI Aql (see
Greiner & Di Stefano 2002), scaled to the LMC distance.
While the early decay of N LMC 2000 was substantially
faster, the optical brightness in Dec. 2000 is consistent
with the nova still being in the plateau phase. For com-
parison purposes also overplotted (with small dots) is the
VSNET light curve of nova V1494 Aql (scaled in bright-
ness) which has very similar t2 and t3 times.
sured column density (note though that the Della Valle
et al. relation refers to the ionized hydrogen mass).
Based on the relation between decay time vs. abso-
lute magnitude (Della Valle & Livio 1995) and our mea-
sured t2 value, the absolute magnitude of Nova LMC 2000
would be MV = −8.74 mag, half a magnitude brighter
than the backwards extrapolation based on the spectral
appearance. This would imply that this extrapolation is
reasonable.
3.3. The supersoft X-ray phase in novae
No supersoft X-ray phase has been found for Nova LMC
2000: while on 25/26 July 2000 the ejected shell was still
too optically thick to allow soft X-rays to penetrate, the
shell had thinned considerably over the next 4 weeks (until
Aug. 28) to allow the detection of supersoft emission, if it
had been present. Thus, the supersoft X-ray phase in Nova
LMC 2000 must have been shorter than 7 weeks, or 6×t2.
In view of the fact that an HST STIS spectrum taken Aug.
19/20, i.e. 9 days before the second XMM-Newton obser-
vation shows a 1150 -- 3120 A flux of 1.6×1038 erg/s (Shore
et al. 2000), the shell burning should either has switched
off during those 9 days, or the effective temperature was
below ∼10 eV in order not to be observable in our XMM-
Newton exposure.
Fig. 4. Correlation between the envelope expansion ve-
locity and the turn-off times of novae (see Tab. 4). For
GQ Mus, Nova Cyg 1992 and V382 Vel two points are
plotted, resembling the temporal limits between observed
supersoft X-ray emission and H burning switch-off. For CI
Aql and U Sco only limits are available. Nova LMC 2000
does not fit into this correlation unless the intrinsic ab-
sorbing column for the supersoft X-ray emission has been
muc larger than estimated in Sect. 2 and 3. If the relation
were true, one would predict that IM Nor with its rather
small expansion velocity would turn into a long-lived (∼3
yrs) supersoft X-ray phase.
The detectability of supersoft X-ray emission from no-
vae depends on several factors, the three most important
ones (probably) being the existence/duration of such a su-
persoft X-ray phase in the first place, the amount of mat-
ter ejected during the thermonuclear runaway and that
blown away later during a wind phase, both of which hide
the supersoft X-ray emission during the early phase after a
nova until the shell has become optically thin. It is there-
fore interesting to ask how these factors relate to other
observable parameters of a nova?
The ignition of the thermonuclear runaway of accreted
hydrogen is primarily determined by the pressure at the
base of the accreted envelope, which in turn depends on
the mass and radius of the white dwarf and the mass of the
envelope (Fujimoto 1982a). It has been argued that the t3-
time primarily depends on the mass of the white dwarf (eq.
12 in Livio 1992). Furthermore, since also the maximum
absolute magnitude of a nova correlates with the mass of
the white dwarf as well as the decline rate, Della Valle
et al. (2002) have recently found a correlation between the
shell mass and the t2-time (note that Shore 2002 has pro-
posed an alternative relation using the same quantities).
If the duration of the supersoft X-ray phase depended on
the amount of mass which is left over after the ejection
6
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
Table 4. Details on novae with detected supersoft X-ray emission (top 5 lines), N LMC 2000, and recent recurrent
novae (RN; last three lines).
Nova Name
t2
t3
(days)
(days)
MWD
(M⊙)(a)
log ∆Mej
(b)
(c)
Vexpa
(km/s)
GQ Mus
N Cyg 1992
N LMC 1995
V382 Vel
V1494 Aql
23.0
16.0
12(e)
4.5
6.6(f )
N LMC 2000
9.0
U Sco (RN)
IM Nor (RN)
CI Aql (RN)
5.0
20.5(b)
30.0
45
42
16(e)
9
16(f )
22
7
47(b)
36
0.87
0.90
1.23
1.17
1.09
1.02
1.20
0.85
0.94
-3.98
-4.02
-4.22
-4.18
-4.13
-4.09
-5.99(g)
-3.99
-3.95
800
2000
900
2700
2800
2200
4500
1150
2400
Porb
(days)
0.0593
0.0812
--
0.1461
0.1346
--
1.230
--
0.6184
log ∆Mign
Limits on supersoft phase
-3.4
-3.8
-4.8
-4.6
-5.3
-5.2
10 yrs
turnoff at days 32 -- 38×t2
>6 yrs and continuing
at 44×t2; none at 60×t2
none <∼28×t2, but at 38 -- 45×t2
less than 5.5×t2
at 4×t2
none at 1.1×t2
less than 14×t2
Refs.
(d)
1 -- 3
4 -- 6
7 -- 9
10
11, 12
13
12, 14
15, 16
(a) According to the t3 − MWD relation (Livio 1992)
(b) According to log ∆Mej = 0.274 × log t2 - 4.355 (Della Valle et al. 2002).
(c) Usually derived for the Hα line. For V382 Vel, lacking an estimate based on Hα, we used the number given for UV lines and
divide by a factor 2 (Shore 2002).
(d) References: (1) Ogelman et al. 1993, (2) Pequignot et al. 1993, (3) Shanley et al. 1995, (4) Krautter et al. 1996, (5) Chochol
et al. 1993, (6) Balman et al. 1998, (7) Della Valle et al. 1995, (8) Orio & Greiner 1999, (9) Orio et al. 1993, (10) Orio et al. 2002,
(11) Venturini et al. 2000, (12) Starrfield et al. 2002, (13) Iijima 2002, (14) Duerbeck et al. 2002, (15) Greiner & Di Stefano 2002,
(16) Kiss et al. 2001.
(d) Determined here from the VSNET light curves.
(g) Instead of the prediction we used the value as measured by Della Valle et al. (2002).
of the shell, i.e. the difference between the mass neces-
sary to ignite the thermonuclear explosion and the mass
of the matter ejected into the expanding shell, it could
be expected that the supersoft X-ray phase would corre-
late with the t2- or t3-time. We have collected the relevant
data (see Tab. 4) for the five novae with detected supersoft
X-ray emission, plus a few cases with stringent limits (in-
cluding the present case Nova LMC 2000). Comparing the
decay times (columns 2 and 3) and the implied envelope
mass (col. 4) according to Della Valle et al. (2002) with the
duration of the supersoft X-ray phase (col. 8) does not
show any obvious correlation, however (though we note
that envelope mass estimates are typically very uncertain).
Thus, the supersoft X-ray phase is seemingly not directly
correlated to the brightness decay rate and/or the mass
of the white dwarf. This also argues against the simple
scenario of Truran & Glasner (1995) and Vanlandingham
et al. (2001): for instance, N LMC 1995 with a rather short
decline time and a white dwarf mass of ∼1.2 M⊙ according
to the t3-MWD relation would have an expected supersoft
X-ray phase of less than 1 yr, but is observed as supersoft
X-ray source for over 6 years now (Orio et al. 2003).
We have also considered the expansion velocities of the
nova shells (col. 6 in Tab. 4), since it is proportional to
the pressure at the base of the accreted shell. There is an
interesting correlation with the duration of the supersoft
X-ray phase in 8 sources except for Nova LMC 2000 which
falls off completely (Fig. 4). Thus, either this correlation
is chance coincidence due to the small number statistics,
or the non-detection the supersoft X-ray phase is caused
by a much higher than estimated absorbing column for
which we have no other evidence. One could imagine that
the (hard) X-ray emission which is seen on 28 August 2000
originates outside the expelled envelope, and therefore the
small absorbing column derived from the fit of that emis-
sion is not representative of the column which blocks the
supersoft emission from the white dwarf. A third alter-
native is that the burning continued much longer, but at
such low temperature (<10 eV) that made it impossible
to be detected by XMM-Newton.
A somewhat surprising correlation is found when plot-
ting the orbital period over the X-ray turn-off time (Fig.
5): systems with a short orbital period have a long H
shell burning period. One possible explanation would be
the higher irradiation of the companion star in short or-
bital period binaries, which in turn may increase the mass
transfer rate to the white dwarf. But even without con-
sidering irradiation, the orbital period is strongly corre-
lated to the mass transfer rate in cataclysmic binaries
(e.g. Patterson 1984), and therefore can be expected to
be a dominant factor in the evolution of a nova through a
supersoft X-ray phase (see below).
A different way to consider the problem of the super-
soft X-ray phase duration is to remember that the H shell
burning may cease before all material is consumed. As has
first been argued by Fujimoto (1982b), burning will cease
when the mass of the shell becomes less than a critical
value Mext, below which the temperature at the bottom
of the envelope is too low for the shell burning to compen-
sate the energy loss from the surface. Thus, the important
factor is not the absolute value of the matter left after the
shell has been expelled, but the difference of this left-over
matter and the critical mass Mext. In evaluating this sug-
gested correlation (Fujimoto 1982b), we have (i) derived
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
7
Aql. The other three systems, V382 Vel, CI Aql and U Sco
happen to fall very near their ∆Mex values, and according
to the suggestion of Fujimoto (1982b) would turn off very
rapidly after the explosion. This is consistent with the
observations.
Of course, there are several limitations to this simple
picture, and additional factors are thought to also influ-
ence the duration of the supersoft X-ray phase, but were
ignored here: (i) Metallicity is certainly an important fac-
tor which determines the appearance of supersoft X-rays.
With Nova LMC 2000 having the same metallicity and
additionally a very similar environment (less than 20 ar-
cmin offset) as the supersoft Nova LMC 1995 (Orio &
Greiner 1999), one could have hoped for a similar X-ray
behaviour. But obviously the limits on the supersoft phase
in Nova LMC 2000 imply a much shorter duration than
the more than 6 years (and continuing) of Nova LMC 1995
(Orio et al. 2003). (ii) It is generally agreed upon that
there is a continuous, long-lasting loss of matter (wind)
after the possible ejection at the time of the thermonu-
clear runaway. Wind-driven mass loss is expected to be
strongly mass dependent, i.e. the radiation pressure grows
for larger white dwarf mass. This would imply a shorter
supersoft X-ray phase for novae with larger white dwarf
masses (e.g. Starrfield et al. 1991, Yungelson et al. 1996).
Unfortunately, no statement can be made concerning
N LMC 2000. It would be interesting to determine the
orbital periods of N LMC 2000 and N LMC 1995 to deter-
mine their location relative to ∆Mex, and thus to test the
above hypothesis that the duration of the supersoft X-ray
phase is determined by the ratio of the left-over mass to
the critical mass ∆Mex for shell burning. If the above re-
lations hold, we would infer orbital periods in the range
of 0.5 -- 1 day for Nova LMC 2000 and 2 hrs for Nova LMC
1995.
Acknowledgments
We are highly indebted to F. Jansen for granting XMM-
Newton Director's discretionary time for these Target of
Opportunity observations of Nova LMC 2000. We are
grateful to the VSNET observers for providing most of the
measurements plotted in Fig. 3. JG thanks I. Bond for the
details of the early photometry obtained within the MOA
project (www.vuw.ac.nz/scps/moa), as well as C. James
and A. Breeveld for the help in the attempt to make use of
the OM grism data. Based on observations obtained with
XMM-Newton, an ESA science mission with instruments
and contributions directly funded by ESA Member States
and NASA.
References
Balman S., Krautter J., Ogelman H., 1998, ApJ 499, 395
Bond I.A., Kilmartin P.M., 2000, IAUC 7457
Chochol D., Hric L., Urban Z., et al., 1993, A&A 277, 103
Duerbeck H.W., Pompei E., 2000, IAUC 7457
Duerbeck H.W., Baptista R., Dutra C.M., Sterken C., 2002,
IAUC 7799
Fig. 5. Correlation between the orbital periods and the
turn-off times of novae (see Tab. 4). Symbols and limits
as in Fig. 4.
Fig. 6. Correlation between the turn-off times of novae
(see Tab. 4) and the ratio of left-over mass to the critical
mass where H shell burning will cease. Symbols and limits
as in Fig. 4.
the accretion rate from the orbital period using the "mas-
ter equation of binary evolution" for the case of rotational
braking (eq. 38 in Patterson 1984), (ii) deduced the white
dwarf mass from the t3 time (eq. 13 in Livio 1992), (iii)
by using the accretion rate and white dwarf masses de-
termined the mass (∆Mign) necessary to ignite the shell
burning depending on the white dwarf mass and accretion
rate (Fig. 7 from Fujimoto 1982b; see column 8 in Tab. 4),
(iv) derived the ejected shell mass from the t2 time (Della
Valle et al. 2002; see column 5 in Tab. 4), (v) compared the
difference of the total shell mass (∆Mign) and the ejected
mass (∆Mej) with the critical mass (∆Mex) which varies
between ∼ 5×10−5 M⊙ for a 0.8 M⊙ white dwarf mass
down to ∼10−6 M⊙ for a 1.2 M⊙ white dwarf mass.
Fig. 6 shows this ratio in dependence of the X-ray turn-
off time for the six novae for which the relevant data (t2,
t3, Porb) are available. We find that three sources have
substantial mass above the critical ∆Mex: GQ Mus has
by far the largest mass excess, consistent with its ∼10 yr
supersoft X-ray phase, as well as N Cyg 1992 and V1494
8
J. Greiner et al.: XMM-Newton observations of Nova LMC 2000
Della Valle M., Livio M., 1995, ApJ 452, 704
Della Valle M., Masetti N., Benetti S., 1995, IAUC 6144
Della Valle M., Pasquini L., Daou D., Williams R.E., 2002,
A&A 390, 155
Fujimoto M.Y., 1982a, ApJ 257, 752
Fujimoto M.Y., 1982b, ApJ 257, 767
Greiner J., DiStefano R., 2002, ApJ 578, L59
Hearnshaw J.B., Yan Tse J., 2000, IAUC 7457
Iijima T., 2002, A&A 387, 1013
Kiss L.L., Thomson J.R., Ogloza W., Furesz G., Sziladi K.,
2001, A&A 366, 858
Krautter J., Ogelman H., Starrfield S., Wichmann R.,
Pfeffermann E., 1996, ApJ 456, 788
Liller W., 2000, IAUC 7453
Livio M., 1992, ApJ 393, 516
Luks T., 1994, RvMA 7, 171
MacDonald J., Fujimoto M.Y., Truran J.W., 1985, ApJ 294,
263
Mason K.O., Breeveld A., Much R. et al. 2001, A&A 365, L36
Ogelman H., Orio M., Krautter J., Starrfield S., 1993, Nat.
361, 331
Orio M., Greiner J., 1999, A&A 344, L13
Orio M., Parmar A.N., Greiner J., et al. 2002, MN 333, L11
Orio M., Hartmann W., Still M., Greiner J., 2003, ApJ (sub-
mitted)
Patterson J., 1984, ApJS 54, 443
Pequignot D., Petitjean P., Boisson C., Krautter J., 1993, A&A
271, 219
Shanley L., Ogelman H., Gallagher J.S., Orio M., Krautter J.,
1995, ApJ 438, L95
Shore S.N., Starrfield S., Bond H.E., Downes R., Hauschildt
P.H., Gehrz R.D., Woodward C.E., Krautter J., Evans
A.N., 2000, IAUC 7486
Shore S.N., 2002,
in Classical Nova Explosions, Eds. M.
Hernanz & J. Jos´e, AIP Conf. Proc. 637, p. 175
Starrfield S., Truran J.W., Sparks W.M., Krautter J., 1991,
in Extreme Ultraviolet Astronomy, ed. R.F. Malina & S.
Bowyer (New York: Pergamon), p. 168
Starrfield S., 2002,
in Classical Nova Explosions, Eds. M.
Hernanz & J. Jos´e, AIP Conf. Proc. 637, p. 89
Struder L., Briel U., Dennerl K., et al. 2001, A&A 365, L18
Truran J.W., 2002, in The Physics of Cataclysmic Cariables
and Related Objects, eds. B.T. Gansicke, K. Beuermann,
K. Reinsch, ASP Conf. 261, p. 576
Truran J.W., Glasner S.A., 1995, in Cataclysmic Variables,
eds. A. Bianchini, M. Della Valle, M. Orio, ASSL 205, 453
Vanlandingham K.M., Schwarz G.J., Shore S.N., Starrfield S.,
2001, AJ 121, 1126
Venturini C., Rudy R.J., Lynch D.K., Mazuk S., Puetter R.C.,
Armstrong T., 2000, IAUC 7490
Yungelson L., Livio M., Truran J.W., et al., 1996, ApJ 466, 890
|
astro-ph/0008041 | 1 | 0008 | 2000-08-02T11:27:54 | Constraints on Cosmological Anisotropy out to z=1 from Supernovae Ia | [
"astro-ph"
] | A combined sample of 79 high and low redshift supernovae Ia (SNe) is used to set constraints on the degree of anisotropy in the Universe out to $z\simeq1$. First we derive the global most probable values of matter density $\Omega_M $, the cosmological constant $\Omega_\Lambda $, and the Hubble constant $H_0$, and find them to be consistent with the published results from the two data sets of Riess et al. 1998 (R98) and Perlmutter et al. 1999 (P99). We then examine the Hubble diagram (HD, i.e., the luminosity-redshift relation) in different directions on the sky by utilising spherical harmonic expansion. In particular, via the analysis of the dipole anisotropy, we divide the sky into the two hemispheres that yield the most discrepant of the three cosmological parameters, and the scatter $\chi^2_{\rm HD}$ in each case. The most discrepant values roughly move along the locus $-4\Omega_M +3 \Omega_{\Lambda} = 1$ (cf. P99), but by no more than $\Delta \approx 2.5$ along this line. For a perfect FRW universe, Monte Carlo realizations that mimic the current set of SNe yield values higher than the measured $\Delta$ in $\sim 1/5$ of the cases. We discuss implications for the validity of the Cosmological Principle, and possible calibration problems in the SNe data sets. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 31 October 2018
(MN LATEX style file v1.4)
Constraints on Cosmological Anisotropy out to z = 1 from
Supernovae Ia
Tsafrir S. Kolatt1 and Ofer Lahav2,1
1 Racah Institute of Physics, The Hebrew University, Jerusalem 91904, Israel
2 Institute of Astronomy, Madingley Rd., CB3 0HA, Cambridge, UK
31 October 2018
ABSTRACT
A combined sample of 79 high and low redshift supernovae Ia (SNe) is used to set
constraints on the degree of anisotropy in the Universe out to z ≃ 1. First we derive the
global most probable values of matter density ΩM , the cosmological constant ΩΛ, and
the Hubble constant H0, and find them to be consistent with the published results from
the two data sets of Riess et al. 1998 (R98) and Perlmutter et al. 1999 (P99). We then
examine the Hubble diagram (HD, i.e., the luminosity-redshift relation) in different
directions on the sky by utilising spherical harmonic expansion. In particular, via the
analysis of the dipole anisotropy, we divide the sky into the two hemispheres that yield
the most discrepant of the three cosmological parameters, and the scatter χ2
HD in each
case. The most discrepant values roughly move along the locus −4ΩM + 3ΩΛ = 1
(cf. P99), but by no more than ∆ ≈ 2.5 along this line. For a perfect FRW universe,
Monte Carlo realizations that mimic the current set of SNe yield values higher than
the measured ∆ in ∼ 1/5 of the cases. We discuss implications for the validity of the
Cosmological Principle, and possible calibration problems in the SNe data sets.
Key words: cosmology: miscellaneous -- cosmology: observations -- cosmology: theory
-- supernovae:general
1
INTRODUCTION
The validity of the Cosmological Principle and the isotropy
it implies gained much credibility in recent years. The small
fluctuations in the CMB (∆T /T ∼ 10−5 on angular scale ∼
10◦) provide the strongest evidence that the universe can be
well approximated by the FRW metric on scales larger than
∼ 1000 h−1 Mpc (e.g., Peebles 1993; Wu, Lahav, & Rees
1999)
On smaller scales (∼ 100 h−1 Mpc) bulk flows of the
order v/c ∼ 10−3 indicate that this isotropy breaks down.
This is also manifested by significant correlation functions of
galaxies and clusters on large scales, and structures like the
Supergalactic Plane and the Great Attractor. The transition
scale to isotropy and homogeneity is still poorly known, and
so is the convergence of the acceleration vector of the Local
Group with respect to the CMB. It is therefore important
to quantify the degree of homogeneity and isotropy as func-
tion of scale. Traditionally this was done by searching for
anisotropy in the distribution of radio sources and back-
ground radiations (Nan & Cai 1996; Evans 1992; Webster ).
Several new methods have been suggested to test isotropy
and homogeneity on redshift scales of z ≈ 0.1 − 5, such as
measurements of in situ CMB temperature (Songaila et al.
c(cid:13) 0000 RAS
1994), the derivation of an independent rest frame from mul-
tiple image lens systems (Kochanek, Kolatt, & Bartelmann
1996), and Faraday rotation signature due to anisotropic
magnetic field (Kronberg 1976; Vall´ee 1990; Nodland & Ral-
ston 1997).
The recent use of SNe as distance indicators (Phillips
1993; Perlmutter et al. 1995; Riess, Press, & Kirshner 1996)
opened a new opportunity for accurate measurements of
anisotropy on cosmological scales that previously have not
been accessible. So far the SNe have been used in order to
constrain the Hubble constant H0 from a nearby sample and
combinations of the matter density ΩM and the cosmologi-
cal constant ΩΛ utilizing SNe at moderate (>∼ 0.3) and high
(∼ 1) redshifts. In the future, SNe samples over a wider
redshift range will provide separate estimates for the two
parameters. It is important to establish the 'universality' of
the measurements of cosmological parameters from SN, as
they are commonly used in joint analysis with other probes
such as the CMB, cluster abundance and peculiar velocities
(Efstathiou 1999; Efstathiou et al. 1999; Bridle et al. 1999;
Bridle et al. ; Tegmark 1999).
Assuming a FRW cosmology, a forth measure can be de-
duced from the 'Hubble diagram' (HD; i.e., the luminosity
-- redshift relation), the χ2
HD measure for the best fit model.
2 Kolatt & Lahav
For a perfect distance indicator this measure indicates devi-
ations of the local potential (i.e., at the location of the SN)
from a pure FRW geometry. However, in the real universe
the deviations can also be due to other sources:
• Intrinsic (astrophysical) scatter in the SN luminosity-
light curve relation.
• Scatter due to the location of the SN within the host
galaxy & the galaxy type.
• Scatter due to dust absorption in the host galaxy, in
the intergalactic medium and in our Galaxy.
• Gravitational lensing along the l.o.s. to the SN (e.g.,
an overdensity along the l.o.s. will enhance the apparent
luminosity of a SN).
Here we explicitly assume that there is no evolution
with redshift in the luminosity-light curve relation. Fortu-
nately, most of the abovementioned effects are on the scale of
the host galaxy, so with large enough sample they would be
averaged out in the calculation of large scale anisotropies.
On the other hand, one should worry about 'anisotropies'
which are simply due to poor matching of different data
sets that sample different portions of the sky, or large angu-
lar effects due to Galactic extinction.
We also note that some of these effects above might be
correlated with other measurements, e.g. if the scatter χ2
HD
detected in SN Hubble diagram is affected by fluctuations
in the potential, then it would be correlated with Integrated
SW (or Rees-Schiama) effect in the CMB fluctuations.
The outline of this paper is as follows, in §2 we present
the unified data set we will be using for the isotropy analy-
sis. The results for cosmological parameters from the entire
sample are presented in §3, the anisotropy measurement is
discussed in §4, and put in a probabilistic context in §5. We
conclude our results in §6.
2 THE UNIFIED DATA SET
An ideal data set of SNe for the goals we have put forward
in the introduction would be a whole-sky homogeneous cov-
erage at various redshifts of SNe. Since such an optimal set
does not exist, the closest data set would be the amalgama-
tion of the two existing, published data sets.
We unify the samples of the Supernova Cosmology
Project (SCP) (Perlmutter et al. 1999) and that of the High-
z Supernova search team (HZS) (Riess et al. 1998). These
include also the data from low redshift of the Cal´an-Tololo
survey (Hamuy et al. 1996). The two groups have different
strategy and different nomenclature for the minimization
problem by which the cosmological parameters are derived.
We have brought the SCP data to comply with the language
of the HZS team
B
B − M f iducial
For each SNe we list its (i) cz in the CMB frame, (ii)
the distance modulus µ = mef f
, (iii) errors for
these two quantities, (iv) Galactic l and b. For the SCP data
the fiducial magnitude, M f iducial
(cf. P99), is obtained by
comparison of the 18 overlapping low redshift SNe as anal-
ysed by the two groups, and equating the distance modulus
of R98 (table 10) to mcorr
B of P99 (table 2). This procedure is
repeated twice, since Riess et al. provide two ways to calcu-
late the distance moduli, "Multi Light Curve Shapes" (LCS)
and "Template". Errors are taken from the tables and a least
B
Figure 1. The sky distribution in Galactic coordinates of the 79
SNe composing the unified sample. The point size is proportional
to (1 + z)−1 of the SNe. Also shown are the (positive) directions
that maximize the ΩM and ΩΛ dipoles using the two methods (cf.
§4), and the CMB dipole direction in the Local Group rest-frame
as measured by COBE.
B
B
B
square minimization is performed in order to obtain the two
best fit values of M f iducial
of P99 (and to recover the Hub-
ble constant dependence they omitted in their calculation).
The two values are M f iducial
+ 5 log H0 = −19.322, −19.453
with χ2/d.o.f of 0.952 and 0.763 for the LCS method and the
TEMPLATE method respectively. The value of M f iducial
is
degenerated with H0, so different H0 calibrations in the two
samples get "absorbed" in the value for M f iducial
. The uni-
fied sample consists of 79 SNe altogether, after the exclusion
of 6 SNe from P99 (taking their "model C" version) and in-
cluding the snap-shot survey from R98 along with 1997ck.
Figure 1 shows the SNe distribution in Galactic coordinates.
The sky coverage is clearly inhomogeneous: the SNe de-
ficiency near the Galactic plane is evident and the clustering
of a few of the observed SNe due to the detection procedure
is clear.
B
3 COSMOLOGICAL PARAMETERS FROM
THE UNIFIED SAMPLE
We follow the statistical analysis as described in R98 and
obtain best values for H0 and probability contours in the
(ΩM , ΩΛ) plane after integration (i.e. marginalization) over
all H0 values and taking into account only physical regions
in that plane.
P99 include the error due to redshift measurements and
peculiar velocities in their magnitude errors, for R98 we
followed their procedure, set σv = 200 km s−1 for SNe of
z < 0.5 and σv = 2500 km s−1 for SNe with z ≥ 0.5, and
translated to the distance modulus, µ, units according to the
assumed cosmological model in the likelihood function. Fig-
ure 2 show the results of the likelihood analysis. The max-
ima of the likelihood functions are obtained for (ΩM , ΩΛ)
values of (0.40, 0.82) and (0.66, 1.36) for the LCS and TEM-
PLATE method respectively. The contour lines correspond
to the 68.3%, 95.4%, and 99.7% confidence levels.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
3
Figure 2. Confidence regions drawn from 79 SNe of the unified sample (see text), using the LCS method (left) and the TEMPLATE
method (right). The best-fit parameters are marked by a star and the quoted best-fit line from P99 (0.8ΩM − 0.6ΩΛ = −0.2) is shown
for reference as the diagonal line across the figure. Non-physical regions in the (ΩM , ΩΛ) plane are excluded.
4 ANISOTROPY MEASUREMENT
The natural expansion for anisotropy detection is in spher-
ical harmonics. The current data are too sparse to allow
analysis in redshift shells.
We expand the four two-dimensional parameter 'fields'
(for ΩM , ΩΛ, H0, and χ2
HD) in spherical harmonics. If the
isotropy assumption is valid we expect deviations from the
average value to be due to noise, and the angular power
spectrum should likewise reflect it. This is unless foreground
effects alter the signal significantly.
The operational way to calculate the expansion coeffi-
cients alm is as follows.
• Build a random distribution of points ("mask") on the
sphere.
• Assign four best-fit parameters to every point based on
minimization over all SNe within angular radius γmin about
this grid point.
• Construct the four residual fields about the global
mean, i.e., δF = (F − hF i)/hF i, where F is ΩM , ΩΛ, H0, or
χ2
HD.
• Expand the δF values as obtained at each grid point in
Spherical Harmonics up to lmax = π/γmin, i.e.
l=lmax
m=+l
δF (θ, φ) =
X
X
am
l Y m
l
.
l=0
m=−l
(1)
In order to include more than 2 SNe in each smoothing bin
(at least two-parameter fit) we obtain γ ≃ 25◦, however
the SNe are not distributed uniformly (cf. Fig. 1) and thus
a minimum angular resolution of ∼ 60◦ is imposed. That
means that for a whole sky coverage the highest significant
multipole, l, is l = 3. There are, though regions that are more
densely covered by SNe data and therefore higher multipoles
can be assessed as well but at a lower signal-to-noise level.
In order to account for the Poisson noise contribution
(and thus to the angular power spectrum in quadrature), we
run a set of 50 random "masks" and repeat the alm calcu-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
lation each time. For each set of alm the power spectrum
coefficients, Cl = (2l + 1)−1 Pl
m=−l alm2 are computed.
The angular power spectrum of the δH0 is an order of
magnitude and more smaller than the noise level (Cl/(2π) ≃
5 × 10−5), in both methods. Figure 3 shows the angular
power spectrum, l(l + 1)Cl/(2π), for the other three fields as
calculated from 50 runs with different random mask points.
The straight weaker lines show the noise level in each field.
The δΩ fields in both methods show signals that exceed the
noise level for the dipole (l = 1) and the quadrupole (l = 2).
Two factors contribute to the noise level, the discrete num-
ber of SNe, and the scatter in the luminosity -- redshift
relation. The former is common to both methods (LCS and
TEMPLATE) and therefore the order of magnitude higher
noise level for the δΩ fields in the LCS method must be
due to the latter. The TEMPLATE method seems to pro-
vide smaller errors and a better match between the two data
sets, as indicated by the lower χ2 level of the fiducial magni-
tude calibration (cf. §2). The δχ2
angular power spectrum
is similar in shape and magnitude in both methods, and lies
an order of magnitude to a factor ∼ 5 above its noise level.
This may indicate there exists a true dipole (or quadrupole)
in this field. From the first multipole of angular power spec-
trum alone, one cannot deduce what is the dipole direction.
We therefore turn to look for the direction by other means.
HD
We search for largest dipole in ΩM , ΩΛ, H0, and χ2
HD.
This has been done in two ways : an actual search over the
sky, dividing the SN population in between two hemispheres,
and equivalently, by solving a maximization problem of the
dipole term with respect to (θ, φ) using the computed alm
coefficients. Both methods yield similar results. We then cal-
culate the confidence regions for each hemisphere separately,
and look for statistical consistency (overlapping contours).
Each test can be applied to each one of the four parameters.
Figure 4 verify the fact that the current SNe data best
constrain a linear combination of the cosmological parame-
ters ΩM , ΩΛ. In all four panels the likelihood maxima move
along the line (P99) −4ΩM + 3ΩΛ = 1, sometimes with
4 Kolatt & Lahav
Figure 3. Angular power spectrum coefficients for the two dimensional fields: δΩM , δΩΛ , and δχ2
field (straight weak lines). Shown are results from the LCS method (right) and the TEMPLATE method (left).
HD
along with the noise level of each
a large distance ∆ between the two maxima for the two
disjoint hemispheres (quoted on the plots). In three cases
the contour levels overlap significantly (see next section for
quantitative evaluation). In the case of the ΩΛ dipole, us-
ing the LCS method, there is no overlap between the 99.7%
confidence levels of the two hemispheres. The discrepancy
stems from the very assymetric distribution of SNe between
the two hemispheres (59 on one versus 20 on the other) and
only one SN (1995at) with z > 0.5 in the 20 SNe sample.
A small error in the distance measurement of this SN, or
a systematic deviation of it from the average LCS relation
may cause such a discrepancy as we demonstrate in the next
section. Elimination of this SN yields a dipole which points
∼ 20◦ away from the original direction, reduced ∆ value
of 2.39, and almost full inclusion of the 99.7% confidence
contour for the larger sample (56 SNe) within the 95.4%
confidence level of the remaining 22 SNe.
Note that a different "mixture" of redshift distribution
to different directions may cause some directions to become
more sensitive to one parameter. E.g, SNe at z ≃ 0.4 − 0.5
are mostly sensitive to the ΩM −ΩΛ combination, as opposed
to higher weight on ΩM as redshift increases (∂∆/∂ΩM >
∂∆/∂ΩΛ). Figure 1 includes the dipole directions (positive)
of Ω in both methods. We observe no coincidence with any
Galactic or CMB direction, moreover not all dipoles point
to the same direction.
The dipole of the δχ2
HD
field points in both methods
toward (l = 80◦, b = −20◦) with hχ2
HDi = 1.00 (1.01) and
largest difference of 0.60 (0.63) for the LCS (TEMPLATE)
method. This dipole direction is suspiciously close to the
Galactic plane.
One worry is that the detected signal is due to the
(mis)match between the two data sets. We therefore re-
peated the computation for each data set separately and
verified that though the noise level increases, the results as
drawn from each one of the data sets are consistent with the
results from the unified set both in magnitude and direction.
5 DEGREE OF ANISOTROPY
The results of the last section, regarding the spherical har-
monic expansion and the various dipole magnitudes, should
now be put in an expected distribution in order to draw
conclusions about the degree of anisotropy.
The hypothesis we are trying to address is that the
SN data do not falsify the FRW geometry as a reliable de-
scription of the z ≃ 1 Universe. This strategy is more effi-
cient than addressing specific anisotropic cosmological mod-
els (C´el´erier 2000a; C´el´erier 2000b). We therefore compute
the probability distribution of the dipole magnitudes within
a FRW universe and confront it with the values obtained for
the real Universe.
A simple two dimensional Kolmogorov-Smirnov test to
falsify the hypothesis that the two contour maps come from
the same underlying distribution of cosmological parame-
ters is inadequate here. Since we have used the maximum
discrepant values in order to obtain the dipole, the two sub-
samples are not randomly selected and therefore can not be
confronted in a KS test.
The probability distribution depends on the actual cos-
mological values and to a lesser extent on the power spec-
trum (via the scatter due to potential fluctuations). For a
self consistency check, the underlying cosmology is taken to
be the "best fit" cosmological model (§3), which we then
sample by Monte-Carlo simulations.
To mimic accurately the SN sample, we use the same
angular locations and redshift values as of the observed sam-
ple. Luminosity distances, magnitude scatter and peculiar
velocities are drawn from Gaussian distributions with the
appropriate observed standard deviation.
The dipole analysis is repeated for 200 mock catalogs
of the SN and the maximal dipole magnitude is calculated
to obtain its distribution for the current sampled SNe.
Table 1 shows the rejection levels of the hypothesis that
the Universe up to z ≃ 1 can be described by a FRW metric.
E.g., using the LCS method and the current sample of SNIa
we expect in 19% of all cases to detect a higher ∆ value for
ΩM dipole, than the observed one.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
5
Figure 4. Confidence regions (same as Fig. 2 drawn from 79 SNe of the unified sample in two hemispheres that maximize the ΩM
dipole (up) and ΩΛ dipole (bottom) using the LCS method (left) and the TEMPLATE method (right). Marked are the number of SNe
in each hemisphere, the best-fit H0, and the distance between peak probabilities (∆).
Table 1
Isotropy rejection levels using ∆
Cosmological
Method
parameter
LCS TEMPLATE
H0
ΩM
ΩΛ
33%
81%
88%
70%
79%
64%
6 DISCUSSION
By the exploitation of the current available SNe data we
have put constraints on the rejection level of the cosmolog-
ical principle validity up to z ≃ 1. A FRW metric is found
to be an adequate description of the Universe. In ∼ 20%
of all realizations of such universes, the dipole signature for
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
anisotropy in the cosmological parameters H0, ΩM and ΩΛ
exceeds the observed one.
Even though such dipole magnitudes are reasonable in
the framework of the FRW model, they may be indicative of
non-cosmological contributions to the angular power spec-
trum. In §1 we listed possible such contributions. If indeed
the Universe up to z ≃ 1 is well represented by a FRW
metric then we can exclude large coherent structures at
z >
∼ 0.3. Such are the structures that may lead to dipole and
quadrupole signatures due to coherent gravitational lensing
magnification/de-magnification and therefore the latter can
be excluded as anisotropy contributors.
That leaves small scale (Galactic) foreground effects to
be the most likely power contributors. The Galactic disk
geometry makes the quadrupole the most significant multi-
pole to be considered, though the solar system offset from
the Galactic center may bring about a dipole contribution
as well. In the current sample the quadrupole term is only
6 Kolatt & Lahav
slightly larger than the noise level and does not allow any
conclusive results. None of the dipole directions for Ω coin-
cides with the Galactic plane and thus they are probably not
correlated with it. Multipoles due to dust extinction may be
affirmed by multiple expansion of the residual colors after
extinction correction (i.e., R98 and P99 appendices).
The one case where two significantly non-overlapping
confidence regions are found for two hemispheres that max-
imize the ΩΛ dipole (LCS), is probably due to a single SN
(1995at) for which the individual errors have been underes-
timated. This case is an exception since the overall χ2 val-
ues for the HD fits are statistically acceptable. Nevertheless,
this case demonstrates the hazard in the draw of conclusions
based on a handful of SNe, for which the error in the error
estimate is uncertain.
In general, the TEMPLATE method provides a better
statistical agreement of the data with an FRW model and
the current SNe data. This is seen from the magnitude match
(cf. §2), tighter constraints from the combined set, smaller
noise levels for all multipoles, and smaller ∆ values for the
Ω dipoles.
We conclude that an isotropic z ∼ 1 universe cannot be
rejected by more than a 1σ level based on the current SNe
data.
ACKNOWLEDGMENTS:
This work was supported by the US-Israel Binational Sci-
ence Foundation, by the Israel Science Foundation, and by
grants from NASA and NSF at UCSC.
REFERENCES
Bridle S. L., Eke V. R., Lahav O., Lasenby A. N., Hobson M. P.,
Cole S., Frenk C. S., Henry J. P., 1999, MNRAS, 310, 565
C´el´erier M., 2000a, A&Ap, 353, 63
C´el´erier M., 2000b, in "Proceedings of the XXXVth Rencon-
tres de Moriond, "Energy Densities in the Universe", astro-
ph/0006273
Efstathiou G., 1999, MNRAS, 310, 842
Efstathiou G., Bridle S. L., Lasenby A. N., Hobson M. P., Ellis
R. S., 1999, MNRAS, 303, L47
Evans T., 1992, Thesis Haverford Coll., PA.
Hamuy M. et al., 1996, AJ, 112, 2408
Kochanek C. S., Kolatt T. S., Bartelmann M., 1996, ApJ, 473,
610
Kronberg P., 1976, in Int. Astron. Union Symp., Vol. 74, p. 367
Nan R., Cai Z., 1996, in IAU Symp. 168: Examining the Big Bang
and Diffuse Background Radiations, Vol. 168, p. 491
Nodland B., Ralston J. P., 1997, Phys.Rev.Lett., 78, 3043
Peebles P. J. E., 1993, Principles of Physical Cosmology. Prince-
ton Univ. Press, Princeton, NJ
Perlmutter S. et al., 1999, ApJ, 517, 565
Perlmutter S. et al., 1995, ApJ, 440, L41
Phillips M. M., 1993, ApJ, 413, L105
Riess A. G. et al., 1998, AJ, 116, 1009
Riess A. G., Press W. H., Kirshner R. P., 1996, ApJ, 473, 88
Songaila A. et al., 1994, nat, 371, 43
Tegmark M., 1999, ApJ, 514, L69
Vall´ee J. P., 1990, ApJ, 360, 1
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0209596 | 1 | 0209 | 2002-09-27T22:46:56 | A Catalog of 157 X-ray Spectra and 84 Spectral Energy Distributions of Blazars observed with BeppoSAX | [
"astro-ph"
] | As a special contribution to the proceedings of the BeppoSAX workshop dedicated to blazar astrophysics we present a catalog of 157 X-ray spectra and the broad-band Spectral Energy Distribution (SED) of 84 blazars observed by BeppoSAX during its first five years of operations. The SEDs have been built by combining BeppoSAX LECS, MECS and PDS data with (mostly) non-simultaneous multi-frequency photometric data, obtained from NED and from other large databases, including the GSC2 and the 2MASS surveys. All BeppoSAX data have been taken from the public archive and have been analysed in a uniform way. For each source we present a SED plot, and for every BeppoSAX observation we give the best fit parameters of the spectral model that best describes the data. The energy where the maximum of the synchrotron power is emitted spans at least six orders of magnitudes ranging from ~ 0.1 eV to over 100 keV. A wide variety of X-ray spectral slopes have been seen depending on whether the synchrotron or inverse Compton component, or both, are present in the X-ray band. The wide energy bandpass of BeppoSAX allowed us to detect, and measure with good accuracy, continuous spectral curvature in many objects whose synchrotron radiation extends to the X-ray band. This convex curvature, which is described by a logarithmic parabola law better than other models, may be the spectral signature of a particle acceleration process that becomes less and less efficient as the particles energy increases. Finally some brief considerations about other statistical properties of the sample are presented. | astro-ph | astro-ph |
Blazar Astrophysics with BeppoSAX and Other Observatories
1in1
A Catalog of 157 X-ray Spectra and 84 Spectral Energy
Distributions of Blazars Observed with BeppoSAX
P.GIOMMI1, M. CAPALBI1, M. FIOCCHI1, E. MEMOLA1, M. PERRI1,2,
S. PIRANOMONTE1,3, S. REBECCHI1 AND E. MASSARO2,4
1ASI Science Data Center, c/o ESA-ESRIN, Frascati, Italy
2Dipartimento di Fisica, Universit´a la Sapienza, P.le A. Moro 2, Roma, Italy
3Dipartimento di Astronomia, Universit´a di Padova, Vicolo dell'Osservatorio 2, Padova, Italy
4Istituto di Astrofisica Spaziale, CNR, Roma, Italy
ABSTRACT. As a special contribution to the proceedings of the BeppoSAX workshop dedi-
cated to blazar astrophysics we present a catalog of 157 X-ray spectra and the broad-band Spec-
tral Energy Distribution (SED) of 84 blazars observed by BeppoSAX during its first five years of
operations. The SEDs have been built by combining BeppoSAX LECS, MECS and PDS data
with (mostly) non-simultaneous multi-frequency photometric data, obtained from NED and
from other large databases, including the GSC2 and the 2MASS surveys. All BeppoSAX data
have been taken from the public archive and have been analysed in a uniform way. For each
source we present a νf (ν) vs ν plot, and for every BeppoSAX observation we give the best fit
parameters of the spectral model that best describes the data. The energy where the maxi-
mum of the synchrotron power is emitted spans at least six orders of magnitudes ranging from
≈ 0.1 eV to over 100 keV. A wide variety of X-ray spectral slopes have been seen depending
on whether the synchrotron or inverse Compton component, or both, are present in the X-ray
band. The wide energy bandpass of BeppoSAX allowed us to detect, and measure with good
accuracy, continuous spectral curvature in many objects whose synchrotron radiation extends
to the X-ray band. This convex curvature, which is described by a logarithmic parabola law
better than other models, may be the spectral signature of a particle acceleration process that
becomes less and less efficient as the particles energy increases. Finally some brief considerations
about other statistical properties of the sample are presented.
1. Introduction
Blazars emission is known to be dominated by strong and highly variable non-thermal
radiation across the entire electromagnetic spectrum. Multi-frequency ground based
observations, combined with data from high energy astronomy satellites, have often
been used to derive the broad-band Spectral Energy Distribution (SED) of blazars, that
is the source intensity as a function of energy, usually represented in the νf (ν) vs ν or
νL(ν) vs ν space. These measurements are consistent with the widely accepted scenario
where blazar emission is due to synchrotron radiation whose power increases with energy
up to a peak value above which it drops sharply. At higher energies the spectrum is
dominated by inverse Compton emission which also smoothly raises until it reaches a
second luminosity peak. The often extreme observational characteristics of blazars are
thought to be the result of the emission from a relativistic jet seen at a very small
angle with respect to the line of sight (e.g. Urry & Padovani 1995), an interpretation
first proposed by Blandford & Rees (1978). According to this scenario the position and
the relative power of the synchrotron and inverse Compton peaks directly depend on
important physical parameters such as the intensity of the magnetic field, the maximum
energy at which electrons can be accelerated, and the relativistic motion and orientation
of the emitting plasma. The synchrotron peak is located at energies ranging from less
than ≈ 0.1 eV (or ν ≈ 1013 Hz) to well over 10 keV (or ν ≈ 1018 Hz) or even 100 keV in
flaring states, demonstrating the existence of a wide variety of physical and geometric
conditions in blazars. For these reasons the Spectral Energy Distribution of blazars has
been and still is the subject of intense research activity. Figure 1 shows the expected
2
Giommi et al.
emission from Synchrotron Self Compton models (SSC) tracing a hypothetical sequence
of blazar SEDs that ranges from LBL sources where the synchrotron peak frequency
(νpeak ) occurs at low energies to HBL objects where νpeak reaches the X-ray band,
and up to the extremely large νpeak energies of the, possibly existing but still unseen,
Ultra High energy peaked BL Lacs (UHBLs). As shown in Figure 1, within the broad-
band energy spectrum of blazars the X-ray region is particularly important since at
these energies a variety of different spectral components can be (and have been) seen.
These include the flat and rising Compton component, the transition between the two
regimes, and the high energy end of the synchrotron spectrum which is produced by very,
sometimes extremely, energetic electrons. These crucial observations, in combination
with other multi-frequency data allow the determination of the overall spectral shape
and therefore the estimation of important physical parameters.
With its very wide X-ray band pass, good sensitivity and spectral capabilities Beppo-
SAX has provided a very important opportunity to study blazars astrophysics, especially
when simultaneous multi-frequency observations could be arranged.
As a special contribution to the proceedings of the BeppoSAX workshop dedicated
to blazar Astrophysics we present the catalog of X-ray spectral fits and broad-band
Spectral Energy Distribution of all the blazars observed with BeppoSAX whose data are
currently public.
2. BeppoSAX Blazars: the sample
A complete description of the BeppoSAX X-ray astronomy satellite is given in Boella et
al. (1997a). The scientific instrumentation is composed of six co-aligned Narrow Field
Instruments (NFI) and two Wide Field Cameras (WFC, Jager et al. 1997) pointing in
opposite directions and perpendicularly to the pointing direction of the NFI. The NFI
include the Low Energy Concentrator Spectrometer, (LECS, Parmar et. al 1997), three
identical Medium Energy Concentrator Spectrometers, (MECS, Boella et al. 1997b),
one of which (MECS1) failed in May 1997, a High Pressure Gas Scintillation Propor-
tional Counter (HPGSPC, Manzo et al. 1997), and a Phoswich Detector System (PDS,
Frontera et al. 1997). Collectively the NFI cover a very wide spectral range (0.1–200
keV) providing an unprecedented opportunity to study the broad band SED of blazars.
During its lifetime BeppoSAX dedicated approximately 15% of its scientific program to
the study of blazars performing about 200 observations of nearly 100 blazars.
We have carried out a systematic analysis of all blazar observations made with the
LECS, MECS and PDS (for sufficiently bright and unconfused sources) instruments
and for which the corresponding data were available in the BeppoSAX public archive
(Giommi & Fiore 1997) at the date of March 2002. As BeppoSAX data are subject to
the usual one year proprietary period, this approximately corresponds to observations
performed in the period July 1996 - March 2001. The blazar class is usually divided in
BL Lacs and Flat Spectrum Radio Quasars (FSRQs), these last also including GigaHertz
Peaked Spectrum (GPS) QSOs.
Our list includes a total of 84 blazars, 58 of which are BL Lacs, 22 are FSRQs, and
4 (1JY 2149−3006, PKS 2126−15, OX 57 and PKS 2243−123) are GPS QSOs. These
sources were observed as part of several BeppoSAX projects and were discovered both
in the radio and the X-ray band, sometimes in surveys with widely different flux limits.
In addition the BeppoSAX selection process clearly favoured objects that were known
to be X-ray bright and possibly detectable by the high energy instruments (PDS). The
sample is therefore highly heterogeneous and cannot be used for detailed statistical anal-
yses requiring strict completeness criteria. The size of the sample and its heterogeneity
nevertheless provide an unprecedented opportunity to explore a very wide portion of
the blazar parameter space.
The lists of BL Lacs and FSRQs considered in this paper are presented in Tables I
Blazar Astrophysics with BeppoSAX and Other Observatories
1in3
Fig. 1. The Spectral Energy Distribution of BL Lacs is shown as a sequence of Synchrotron Self
Compton spectra peaking at different energies. Objects whose synchrotron component peaks
at low energy are called LBLs, the maximum of their synchrotron power output occurs in the
IR-Optical band, while High energy peaked BL Lacs (HBLs) peak in the UV or X-ray band.
It is not known how far the sequence of νpeak goes on. If it continues to very high energies it
could be that in some very extreme objects (UHBLs) the synchrotron component might even
reach the gamma-ray band. Note that for the same peak luminosity, the radio power decreases
by orders of magnitudes in going form LBLs to HBLs and possibly to UHBLs.
and II, respectively. Column 1 gives the name of the object, columns 2 and 3 give the
Right Ascension and Declination for the equinox 2000.0, column 4 gives the redshift
when available, column 5 gives the number of BeppoSAX observations available in the
public archive, and column 6 gives the references to previous BeppoSAX publications.
In general, sources brighter than ≈ 5 × 10−12 erg cm−2 s−1 can be detected in all
three BeppoSAX NFI instruments considered in this paper (LECS, MECS, and PDS),
fainter sources can only be detected in the imaging instruments (LECS and MECS) up
to 10 keV. A total of 157 spectra have been analysed.
3. Data analysis
All LECS, MECS and PDS cleaned and calibrated spectral data (pha files) have been
taken from the BeppoSAX public archive (Giommi & Fiore 1997). The standard Beppo-
SAX procedure that generated these archival spectra used photons located within a
circular region centered on the source with a typical radius of 4 arcminutes for the
MECS data and of 6 arcminutes for the LECS data. For the case of weak sources,
4
Giommi et al.
however, this procedure used a radius of 4 and sometimes even 2 arcminutes to extract
the LECS spectra, a value that turned out to be too small to ensure a good signal-to-
noise ratio. All LECS spectral files that were originally produced using a 2 arcminutes
radius have been re-generated by us using a more adequate 4 arcminutes radius. In a
few cases non-standard extraction radii had to be used to avoid contamination from a
nearby source. Background subtraction was carried out using the standard background
files available from the BeppoSAX archive. Since these background data were taken
observing "blank" areas of the sky (that is fields not including a detectable target) with
low Galactic absorption (NH ), the low energy counts of X-ray sources located in regions
of high NH may be underestimated. We have therefore ignored data below 0.3 keV.
Finally, spectral channels were binned in such a way that each channel includes at least
20 counts.
All data were analysed in a homogeneous way following the recommendations given
in the BeppoSAX handbook for NFI spectral analysis (Fiore et al. 1999).
We have combined LECS data covering the range 0.3–2.0 keV, MECS data between
2–10 keV, and, for bright and unconfused sources, PDS data above 12 keV and up to
the maximum energy at which these sources could be detected.
3.1. Spectral models
Spectral fits have been done with the XSPEC 11.0 package using one of the following
spectral models
1. single power law
dN/dE = ke−σ(E)NH E−γ
2. sum of two power laws
dN/dE = e−σ(E)NH (kE−γ + k1E−γ1)
3. broken power law
dN/dE =( ke−σ(E)NH E−γ
kEγ1−γ
breake−σ(E)NH E−γ1
if E < Ebreak
if E ≥ Ebreak
4. logarithmic parabola
dN/dE = ke−σ(E)NH E(−γ+βLog(E))
in this last case γ is the photon spectral index at 1 keV and β is the coefficient of
the quadratic term in the logarithmic parabola (Log(dN/dE) = log(k) − γLog(E) +
β(Log(E))2), which is proportional to the slope change in an energy decade. This spec-
tral law was first used by Landau et al. (1986) who were able to obtain very good fits
to the wide band spectra (mm to optical-UV) of a number of BL Lacs. More recently,
Massaro et al. (2000) and Cusumano et al. (2001) successfully applied this model to
fit the X-ray spectra of young pulsars and developed the XSPEC routine used in the
present work. With respect to other curved spectra, the log-parabola has the advantage
of describing well the spectral curvature that is often seen in the broad band spectrum
of blazars with only three parameters. Furthermore, as discussed by Massaro (2002),
it is possible to show that a log-parabolic spectrum is naturally obtained by a statis-
tical acceleration mechanism where the probability for a particle to remain inside the
acceleration region is assumed to decrease as the particles energy grows.
The amount of NH in the low energy absorption term (e−σ(E)NH ) was always fixed
to the Galactic value as estimated from the 21 cm measurements of Dickey & Lockman
(1990).
Blazar Astrophysics with BeppoSAX and Other Observatories
1in5
The LECS-MECS and MECS-PDS relative normalizations were treated as free pa-
rameters but constrained to be within 0.6 and 1.1 and 0.8-1.0, respectively. Typi-
cal values are 0.7 for LECS/MECS and 0.9 for MECS/PDS, as expected from the
BeppoSAX NFI intercalibration (see Fiore et al. 1999 for details).
3.2. Results of the spectral fitting
The results of our spectral fits are reported in Tables III-IV, where, for each source, we
give the best fit parameters for the spectral model that gives the lowest χ2. For sources
that were observed more than once only one model was used choosing the one that gives
the lowest χ2 in most of the observations.
Column 1 gives the source name built with the J2000.0 coordinates (for reasons of
space we use here a IAU like naming convention rather than the full name), column 2
gives the observation date, columns 3, 4, 5, 6 and 7 give the model used and the best
fit parameters (with one sigma errors), column 8 gives the reduced χ2 and the number
of degrees of freedom, column 9 gives the 2–10 keV flux in units of erg cm−2 s−1 .
To keep the amount of work manageable we have neglected the effects of rapid vari-
ability always integrating the spectral data over the full observation. The high χ2 values
that resulted from the analysis of some rapidly variable sources are due to significant
spectral changes that usually accompany large intesity variations (e.g. Giommi et al.
1990, Tanihata et al. 2001). The best estimates of the β parameter in model 4 are
mostly negative and between −0.5 and −0.15; the few positive values attest the pres-
ence of an upward spectral curvature due to presence of both the synchrotron and the
Compton component in the BeppoSAX band.
As all of the data used in this work are publicly available, a large fraction of the
BeppoSAX spectra has already been analysed and published by the original investigators.
We have therefore compared the results of our fits with published results and we have
found that, when the spectral model used is the same, our best fit parameters agree well
with the published values.
3.3. Spectral Energy Distributions
In order to build the broad band (radio to hard X-ray) Spectral Energy Distributions
we have first converted all BeppoSAX spectra into flux using the best model and best fit
parameters listed in Tables III-IV. We have then de-reddened the soft X-ray fluxes using
the cross sections of Morrison & McCammon (1983) setting the amount of absorbing
material (NH ) equal to the Galactic value, consistently with the fitting procedure. We
have then combined these X-ray fluxes with (mostly non-simultaneous) multi-frequency
photometric measurements available from NED or from specific BeppoSAX publications.
We have also added infrared and optical fluxes converting magnitudes from the blazars
counterparts in the 2MASS and GSC2 catalogs (when available) using the prescriptions
given in Cardelli et al. (1989). In several nearby objects (e.g. Mkn 421, Mkn 501, 3C 66A,
PKS 0548−322) the optical emission is dominated by the host galaxy and therefore our
fluxes should be considered as an upper limit to the non-thermal nuclear emission.
All SEDs are plotted in Figures 2a through 2o for BL Lacs and in Figures 3a through
3g for FSRQs. Fluxes have not been corrected for redshift and therefore are in the
observer's frame.
All the SEDs, including the data files, and spectral fitting details are also available
on-line from the ASI Science Data Center web server at the following address
http://www.asdc.asi.it/blazars/
6
Giommi et al.
BL Lacs available in the BeppoSAX public archive
TABLE I
Blazar name
Ra(J2000.0) Dec(J2000.0) Redshift
No. of
observ.
Ref.
1ES 0033+595
PKS 0048−097
1ES 0120+340
RGB J0136+39
1ES 0145+138
MS 0158.5+0019
3C 66A
AO 0235+164
MS 0317.0+1834
1H 0323+022
1ES 0347−121
1H 0414+009
1ES 0502+675
1H 0506−039
PKS 0521−365
PKS 0537−441
PKS 0548−322
S5 0716+714
MS 0737.9+7441
OJ 287
B2 0912+29
1ES 0927+500
1ES 1028+511
RGB J1037+57
RGB J1058+56
1H 1100−230
Mkn 421
1RXS J111706.3+20141
EXO 1118.0+4228
Mkn 180
PKS 1144−379
1RXS J121158.1+22423
EXO 1215.3+3022
1H 1219+301
ON 231
1RXS J123511.1−14033
1ES 1255+244
MS 1312.1−4221
EXO 1415.6+2557
PG 1418+546
1H 1430+423
MS 1458.8+2249
1H 1515+660
PKS 1519−273
1ES 1533+535
1ES 1544+820
1ES 1553+113
Mkn 501
00 35 52.7
00 50 41.3
01 23 08.6
01 36 32.7
01 48 29.7
02 01 06.1
02 22 39.6
02 38 38.8
03 19 51.9
03 26 13.8
03 49 23.2
04 16 52.3
05 07 56.2
05 09 38.0
05 22 57.7
05 38 50.3
05 50 41.9
07 21 53.4
07 44 05.4
08 54 48.9
09 15 52.2
09 30 37.6
10 31 18.5
10 37 44.3
10 58 37.7
11 03 37.7
11 04 27.2
11 17 06.3
11 20 48.0
11 36 26.4
11 47 01.3
12 11 58.0
12 17 52.1
12 21 22.0
12 21 31.7
12 35 11.1
12 57 31.8
13 15 03.4
14 17 56.2
14 19 46.6
14 28 32.4
15 01 01.8
15 17 47.4
15 22 37.6
15 35 00.7
15 40 15.6
15 55 43.1
16 53 52.2
59 50 03.8
−09 29 04.9
34 20 48.8
39 06 00.0
14 02 17.8
00 34 00.1
43 02 08.1
16 36 59.0
18 45 33.8
02 25 14.8
−11 59 26.8
01 05 24.0
67 37 23.8
−04 00 46.0
−36 27 02.8
−44 05 08.8
−32 16 10.9
71 20 35.8
74 33 57.9
20 06 30.9
29 33 23.0
49 50 26.1
50 53 34.0
57 11 54.9
56 28 10.9
−23 29 30.8
38 12 32.0
20 14 09.9
42 12 11.8
70 09 28.0
−38 12 11.1
22 42 36.0
30 07 00.1
30 10 36.8
28 13 58.0
−14 03 32.0
24 12 39.9
−42 36 50.0
25 43 55.9
54 23 15.0
42 40 23.8
22 38 03.8
65 25 23.1
−27 30 11.1
53 20 36.9
81 55 05.8
11 11 21.1
39 45 37.0
−
−
0.272
−
0.125
0.299
0.444
0.94
0.19
0.147
0.188
0.287
0.314
0.304
0.055
0.896
0.069
−
0.315
0.306
−
0.188
0.361
−
−
0.186
0.031
0.137
0.124
0.046
1.048
0.77
0.237
0.13
0.102
0.406
0.141
0.105
0.237
0.152
0.13
0.235
0.702
−
0.89
−
0.36
0.033
1
2
2
1
1
1
2
1
1
1
1
1
1
1
1
1
3
3
1
1
1
1
1
1
2
2
15
1
1
1
1
3
2
1
2
3
1
1
3
3
1
1
1
1
1
1
1
11
a
b
a
c
d
e
d
c
d
d
d
c
m
n
a
g
d
b,f
b
c
c
d
h,i,j
c
d
b
k
c
e
f
d
c
c
c
l
a Costamante et al. 2001, b Padovani et al. 2001, c Beckmann et al. 2002, d Wolter et al. 1998,
e Padovani et al. 2002a, f Massaro et al. 2002, g Giommi et al. 1999, h Fossati et al. 2000a,
i Fossati et al. 2000b, j Malizia et al. 2000, k Tagliaferri et al. 2000, lPian et al. 1998,
mTavecchio et al. 2002, nPian et al. 2002
Blazar Astrophysics with BeppoSAX and Other Observatories
1in7
BL Lacs available in the BeppoSAX public archive - Continued
TABLE I
Blazar name
Ra(J2000.0) Dec(J2000.0) Redshift
No. of
observ.
Ref.
1ES 1741+196
S5 1803+784
3C 371
S4 1823+568
1ES 1959+650
PKS 2005−489
PKS 2155−304
BL Lacertae
1ES 2344+514
1H 2354−315
17 43 57.5
18 00 45.6
18 06 50.7
18 24 07.1
19 59 59.9
20 09 25.3
21 58 51.9
22 02 43.2
23 47 04.7
23 59 07.8
19 35 09.9
78 28 04.0
69 49 27.8
56 51 01.0
65 08 54.9
−48 49 54.1
−30 13 31.0
42 16 40.0
51 42 17.9
−30 37 39.0
0.083
0.684
0.051
0.664
0.048
0.071
0.116
0.069
0.044
0.165
1
1
1
1
3
2
3
5
7
1
a
b
b
c
d
c
e,f,g
c,h
i
a Siebert et al. 1999, b Padovani et al. 2002a, c Padovani et al. 2001, d Beckmann et al. 2002,
e Giommi et al. 1998, f Chiappetti et al. 1999, g Zhang et al. 1999, h Ravasio et al. 2002,
i Giommi et al. 2000
Flat Spectrum Radio Quasars available in the BeppoSAX public archive
TABLE II
Blazar name
Ra(J2000.0) Dec(J2000.0) Redshift
No. of
observ.
Ref.
PKS 0208−512
NRAO 140
PKS 0523−33
OG 147
WGA J0546.6−6415
1Jy 0836+710
RGB J0909+039
1Jy 1127−145
3C 273
3C 279
B3 1428+422
GB 1508+5714
PKS 1510−08
1ES 1627+402
3C 345
V 396 HER
4C 62.29
S5 2116+81
PKS 2126−15
OX 57
1Jy 2149−306
PKS 2223+21
3C 446
CTA 102
PKS 2243−123
3C 454.3
02 10 46.1
03 36 30.0
05 25 06.1
05 30 56.3
05 46 42.4
08 41 24.4
09 09 15.8
11 30 07.0
12 29 06.6
12 56 11.1
14 30 23.5
15 10 02.8
15 12 50.4
16 29 01.3
16 42 58.7
17 22 41.3
17 46 13.9
21 14 01.6
21 29 12.0
21 36 38.4
21 51 55.3
22 25 38.0
22 25 47.2
22 32 36.4
22 46 18.1
22 53 57.6
−51 01 01.9
32 18 29.8
−33 43 05.0
13 31 54.8
−64 15 21.9
70 53 40.9
03 54 42.1
−14 49 27.1
02 03 07.9
−05 47 21.1
42 04 36.1
57 02 47.0
−09 05 59.9
40 07 58.0
39 48 36.0
24 36 19.0
62 26 53.8
82 04 46.9
−15 38 41.9
00 41 53.8
−30 27 53.9
21 18 06.1
−04 57 01.0
11 43 50.8
−12 06 51.8
16 08 53.1
1.003
1.259
4.401
2.06
0.323
2.19
3.2
1.187
0.158
0.538
4.715
4.301
0.361
0.272
0.594
0.175
3.886
0.086
3.266
1.936
2.34
1.949
1.404
1.037
0.63
0.86
1
1
1
8
1
1
1
1
9
5
1
2
1
1
1
1
1
2
1
1
1
1
1
5
1
1
a
b
c
d
e
f
g
h
i
f
e
d
a
d
f
d
j
a,f
e
a
a
a Tavecchio et al. 2002, b Fabian et al. 2001a, c Ghisellini et al. 1999, d Padovani et al. 2002b,
e Tavecchio et al. 2000, f Costantini et al. 1999, g Mineo et al. 2000, h Ballo et al. 2002,
i Fabian et al. 2001b, j Elvis et al. 2000
8
Giommi et al.
Results of spectral fits for the sample of BL Lacertae Objects
TABLE III
Blazar name
Observ. model(a)
0035+5950
0050−0929
0123+3420
0136+3906
0148+1402
0201+0034
0222+4302
0238+1636
0319+1845
0326+0225
0349−1159
0416+0105
0507+6737
0509−0400
0522−3627
0538−4405
0550−3216
0721+7120
0744+7433
0854+2006
0915+2933
0930+4950
1031+5053
1037+5711
1058+5628
1103−2329
1104+3812
1117+2014
1120+4212
1136+7009
1147−3812
1211+2242
1217+3007
1221+3010
1221+2813
date
18-12-99
19-12-97
03-01-99
02-02-99
09-01-01
30-12-97
16-08-96
31-01-99
28-01-99
15-01-97
20-01-98
10-01-97
21-09-96
06-10-96
11-02-99
02-10-98
28-11-98
20-02-99
07-04-99
14-11-96
07-11-98
30-10-00
29-10-96
24-11-97
14-11-97
25-11-98
01-05-97
02-12-98
21-05-98
23-11-98
04-01-97
19-06-98
29-04-97
30-04-97
01-05-97
02-05-97
03-05-97
04-05-97
05-05-97
21-04-98
23-04-98
22-06-98
04-05-99
26-04-00
28-04-00
30-04-00
09-05-00
13-12-99
01-05-97
10-12-96
10-01-97
27-12-99
23-12-98
12-07-99
11-05-98
11-06-98
γ
1.42 ± 0.08
1.9 ± 0.2
1.8 ± 0.1
1.9 ± 0.2
2.0 ± 0.1
2.2 ± 0.4
2.8 ± 0.3
2.2 ± 0.1
1.9 ± 0.2
1.5 ± 0.3
2.2 ± 0.1
1.8 ± 0.2
2.6 ± 0.06
2.31 ± 0.04
2.00 ± 0.07
1.75 ± 0.04
1.75 ± 0.07
1.53 ± 0.13
1.74 ± 0.15
2.3 ± 0.1
2.1 ± 0.2
2.6 ± 0.1
2.62 ± 0.15
1.62 ± 0.12
2.36 ± 0.09
1.7 ± 0.1
2.40 ± 0.05
2.5 ± 0.1
2.7 ± 0.3
2.8 ± 0.1
2.07 ± 0.03
2.25 ± 0.02
2.23 ± 0.03
2.21 ± 0.03
2.19 ± 0.03
2.16 ± 0.04
2.24 ± 0.06
2.49 ± 0.07
2.29 ± 0.04
2.12 ± 0.01
1.99 ± 0.01
2.03 ± 0.02
2.30 ± 0.01
1.75 ± 0.01
1.72 ± 0.01
1.72 ± 0.01
1.82 ± 0.01
2.34 ± 0.09
2.7 ± 0.1
2.1 ± 0.2
1.6 ± 0.2
1.6 ± 0.3
3.0 ± 2.1
2.21 ± 0.08
2.8 ± 0.1
2.9 ± 0.3
4
1
4
4
4
1
4
1
1
4
1
4
1
1
1
1
1
4
4
3
3
3
1
1
1
4
1
1
1
1
1
1
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
1
4
1
4
2
4
2
2
Best fit parameters
β
γ1
−0.44 ± 0.06
Ebreak
−0.2 ± 0.1
−0.4 ± 0.2
−0.5 ± 0.1
0.5 ± 0.3
−0.5 ± 0.3
−0.4 ± 0.2
−0.49 ± 0.11
−0.45 ± 0.14
1.8 ± 0.2
1.7 ± 0.2
1.7 ± 0.1
3.7 ± 0.6
2.9 ± 0.9
3.0 ± 0.3
−0.4 ± 0.1
−0.46 ± 0.03
−0.51 ± 0.03
−0.47 ± 0.03
−0.51 ± 0.04
−0.52 ± 0.06
−0.47 ± 0.07
−0.54 ± 0.04
−0.47 ± 0.01
−0.43 ± 0.01
−0.39 ± 0.02
−0.53 ± 0.01
−0.26 ± 0.01
−0.26 ± 0.01
−0.32 ± 0.01
−0.23 ± 0.01
−0.6 ± 0.1
−0.4 ± 0.2
−0.5 ± 0.3
−0.31 ± 0.07
1.0 ± 1.5
0.6 ± 0.3
1.1 ± 0.4
χ2/d.o.f.
(d.o.f.)
0.95(85)
0.52(9)
1.1(45)
0.9(45)
1.39(44)
0.57(8)
1.21(22)
0.75(24)
0.9(22)
0.74(39)
1.3(22)
0.92(42)
0.75(34)
1.34(37)
1.20(37)
0.83(54)
0.68(42)
0.96(56)
0.82(43)
1.11(41)
1.46(36)
0.99(48)
1.32(18)
0.99(17)
1.36(42)
1.08(42)
1.05(40)
0.98(22)
0.68(14)
0.91(16)
1.10(242)
1.32(85)
1.40(108)
0.76(102)
1.03(102)
0.91(104)
0.98(89)
0.97(84)
1.17(88)
1.10(140)
1.50(136)
1.27(123)
1.44(137)
2.72(130)
1.67(130)
1.72(130)
3.05(130)
1.89(41)
0.73(19)
0.85(36)
1.02(17)
1.24(41)
1.09(17)
0.83(88)
0.97(60)
1.19(46)
flux(b)
5.83
0.13
1.71
1.31
1.05
0.05
0.3
0.20
0.08
0.7
0.3
0.6
0.85
1.94
0.56
0.88
0.24
2.28
1.52
0.14
0.26
0.33
0.15
0.21
0.26
0.67
1.03
0.09
0.03
0.17
3.81
2.56
8.5(c)
8.3
10.0
12.3
7.1
4.1
6.3
17.8
31.0(c)
23.9
11.1
61.9(c)
66.8(c)
59.8(c)
49.5(c)
0.61
0.29
0.51
0.09
0.24
0.07
1.49
0.42
0.33
(a) 1: single power law, 2: double power law, 3: broken power law, 4: logarithmic parabola model
(b) X-ray flux in the 2–10 keV band in units of 10−11 erg cm−2 s−1
(c) The source flux strongly varied during the observation
Blazar Astrophysics with BeppoSAX and Other Observatories
1in9
Results of spectral fits for the sample of BL Lacertae Objects - Continued
TABLE III
Blazar name
Observ.
model(a)
1235−1403
1257+2412
1315−4236
1417+2543
1419+5423
1428+4240
1501+2238
1517+6525
1522−2730
1535+5320
1540+8155
1555+1111
1653+3945
1743+1935
1800+7828
1806+6949
1824+5651
1959+6508
2009−4849
2158−3013
2202+4216
2347+5142
2359−3037
date
27-06-99
16-07-99
20-06-98
21-02-97
13-07-00
23-07-00
27-07-00
12-02-99
03-03-00
26-03-00
08-02-99
19-02-01
05-03-97
01-02-98
13-02-99
13-02-99
05-02-98
07-04-97
11-04-97
16-04-97
28-04-98
29-04-98
01-05-98
20-06-98
29-06-98
16-07-98
25-07-98
10-06-99
26-09-98
28-09-98
22-09-98
11-10-97
04-05-97
29-09-96
01-11-98
20-11-96
22-11-97
04-11-99
08-11-97
05-06-99
05-12-99
26-07-00
31-10-00
03-12-96
04-12-96
05-12-96
07-12-96
11-12-96
26-06-98
03-12-99
21-06-98
γ
1.6 ± 0.7
2.8 ± 0.5
2.14 ± 0.08
2.2 ± 0.1
1.7 ± 0.1
1.5 ± 0.3
1.8 ± 0.1
1.7 ± 0.2
1.3 ± 0.2
1.9 ± 0.2
1.90 ± 0.02
2.7 ± 0.1
2.51 ± 0.06
2.0 ± 0.2
2.42 ± 0.09
2.6 ± 0.1
2.2 ± 0.2
1.69 ± 0.01
1.67 ± 0.01
1.43 ± 0.01
1.64 ± 0.02
1.62 ± 0.02
1.71 ± 0.02
1.81 ± 0.02
1.70 ± 0.02
1.73 ± 0.03
1.78 ± 0.02
2.16 ± 0.01
2.08 ± 0.04
1.51 ± 0.08
2.5 ± 0.5
1.9 ± 0.2
2.1 ± 0.2
2.3 ± 0.1
2.05 ± 0.02
2.38 ± 0.02
2.28 ± 0.02
2.57 ± 0.02
1.84 ± 0.06
2.4 ± 0.2
1.56 ± 0.03
1.84 ± 0.06
2.63 ± 0.03
1.7 ± 0.1
1.8 ± 0.1
2.0 ± 0.1
1.62 ± 0.07
1.7 ± 0.1
2.34 ± 0.03
2.0 ± 0.1
1.82 ± 0.07
4
4
1
1
4
4
4
1
1
1
1
1
1
1
1
1
4
4
4
4
4
4
4
4
4
4
4
4
1
1
4
1
4
4
4
4
4
4
1
2
1
1
1
4
4
4
4
4
1
4
4
Best fit parameters
β
γ1
Ebreak
−0.8 ± 0.7
0.4 ± 0.5
−0.5 ± 0.1
−0.8 ± 0.3
−0.3 ± 0.1
−0.6 ± 0.1
−0.19 ± 0.01
−0.11 ± 0.01
−0.14 ± 0.01
−0.17 ± 0.02
−0.18 ± 0.01
−0.26 ± 0.02
−0.22 ± 0.02
−0.22 ± 0.02
−0.33 ± 0.02
−0.27 ± 0.02
−0.24 ± 0.01
0.7 ± 0.4
−0.4 ± 0.2
−0.05 ± 0.11
−0.12 ± 0.02
−0.27 ± 0.02
−0.46 ± 0.02
−0.30 ± 0.02
−0.3 ± 0.1
−0.15 ± 0.09
−0.14 ± 0.09
−0.15 ± 0.07
−0.4 ± 0.1
−0.3 ± 0.1
−0.23 ± 0.06
1.2 ± 0.4
χ2/d.o.f.
(d.o.f.)
0.63(15)
0.40(18)
0.94(36)
0.94(20)
1.03(49)
0.96(42)
1.11(49)
1.28(15)
0.94(19)
0.93(15)
1.04(89)
1.01(22)
1.20(38)
0.73(17)
0.88(40)
0.71(23)
1.40(45)
1.39(122)
1.38(139)
1.06(139)
1.11(122)
1.18(122)
1.06(122)
1.10(122)
1.07(122)
1.07(122)
1.06(122)
1.37(122)
0.97(41)
0.90(40)
0.62(19)
0.85(12)
0.89(40)
1.23(75)
1.00(124)
1.27(121)
1.22(102)
1.30(95)
1.02(46)
1.07(59)
0.97(53)
0.65(41)
0.98(96)
0.77(40)
1.23(45)
1.03(45)
0.99(87)
0.85(43)
0.79(45)
1.08(83)
1.36(91)
flux(b)
0.19
0.16
1.12
0.44
1.27
0.85
0.93
0.06
0.11
0.08
2.0
0.07
0.99
0.06
0.26
0.15
1.30
21.5
23.9
52.4
17.6
20.9
14.7
4.95
10.3
9.0
9.3
3.0
0.69
0.22
0.19
0.10
1.35
6.01
17.6
5.59
8.28
2.47
1.11
0.66
1.26
0.59
2.03
1.8
2.0
2.9
3.7
2.2
0.8
0.9
2.9
(a) 1: single power law, 2: double power law, 3: broken power law, 4: logarithmic parabola model
(b) X-ray flux in the 2–10 keV band in units of 10−11 erg cm−2 s−1
10
Giommi et al.
Results of spectral fits for Flat Spectrum Radio Quasars
TABLE IV
Blazar
0210−5101
0336+3218
0525−3343
0530+1331
0546−6415
0841+7053
0909+0354
1130−1449
1229+0203
1256−0547
1430+4204
1510+5702
1512−0905
1629+4007
1642+3948
1722+2436
1746+6226
2114+8204
2129−1538
2136+0041
2151−3027
2225+2118
2225−0457
2232+1143
2246−1206
2253+1608
Observ.
date
14-01-01
05-08-99
27-02-00
21-02-97
22-02-97
27-02-97
01-03-97
03-03-97
04-03-97
06-03-97
11-03-97
01-10-98
27-05-98
27-11-97
04-06-99
18-07-96
13-01-97
15-01-97
17-01-97
22-01-97
24-06-98
09-01-00
13-06-00
12-06-01
13-01-97
15-01-97
21-01-97
23-01-97
19-06-98
04-02-99
29-03-97
01-02-98
03-08-98
11-08-99
19-02-99
13-02-00
29-03-97
29-04-98
12-10-98
24-05-99
25-11-00
31-10-97
12-11-97
10-11-97
11-11-97
13-11-97
16-11-97
18-11-97
21-11-97
18-11-98
05-06-00
model(a)
Best fit parameters
γ1
β
Ebreak
0.4 ± 0.2
1.58 ± 0.01
1.58 ± 0.03
1.57 ± 0.03
1.62 ± 0.01
1.56 ± 0.04
1.61 ± 0.02
1.63 ± 0.02
1.62 ± 0.02
1.58 ± 0.03
1.0 ± 0.3
1.0 ± 0.3
1.0 ± 0.4
1.0 ± 0.2
1.6 ± 44
1.0 ± 0.3
1.18 ± 0.08
0.8 ± 0.6
1.0 ± 0.3
−0.3 ± 0.1
−0.2 ± 0.1
−0.42 ± 0.07
γ
1.67 ± 0.06
1.65 ± 0.05
1.6 ± 0.1
1.3 ± 0.1
1.4 ± 0.1
1.5 ± 0.2
1.3 ± 0.1
1.3 ± 0.1
1.5 ± 0.1
1.5 ± 0.1
1.4 ± 0.2
2.2 ± 0.3
1.30 ± 0.02
1.26 ± 0.12
1.42 ± 0.04
1.2 ± 0.3
1.9 ± 0.4
1.3 ± 0.4
1.3 ± 0.2
1.6 ± 0.3
1.8 ± 0.3
2.2 ± 0.3
1.7 ± 0.3
1.9 ± 0.4
1.65 ± 0.06
1.66 ± 0.05
1.76 ± 0.07
1.69 ± 0.05
2.25 ± 0.02
1.38 ± 0.05
1.3 ± 0.5
1.2 ± 0.4
1.36 ± 0.05
2.5 ± 0.1
1.59 ± 0.06
1.7 ± 0.1
1.60 ± 0.09
1.4 ± 0.1
1.5 ± 0.2
1.08 ± 0.09
1.57 ± 0.06
1.39 ± 0.04
1.3 ± 0.1
1.9 ± 0.2
1.51 ± 0.06
1.61 ± 0.06
1.55 ± 0.06
1.53 ± 0.08
1.58 ± 0.07
1.6 ± 0.1
1.40 ± 0.04
1
1
4
1
1
1
1
1
1
1
1
4
1
1
1
3
3
3
3
3
3
3
3
3
1
1
1
1
1
1
1
1
1
1
1
1
1
4
4
4
1
1
1
1
1
1
1
1
1
1
1
χ2/d.o.f.
(d.o.f.)
1.81(41)
1.16(49)
1.1(17)
1.35(17)
1.2(16)
0.91(17)
0.84(18)
1.06(19)
0.72(17)
1.13(16)
0.69(14)
0.98(38)
1.04(101)
0.83(17)
0.95(65)
1.36(116)
0.87(112)
1.34(112)
1.14(112)
1.09(99)
1.21(126)
1.46(125)
1.03(110)
0.87(112)
0.92(43)
1.00(42)
1.04(41)
1.47(42)
1.32(85)
0.98(40)
0.47(1)
1.99(4)
0.77(41)
1.27(21)
1.29(57)
1.08(17)
0.44(39)
1.29(45)
1.12(39)
1.08(95)
0.85(36)
0.69(46)
1.28(18)
1.53(16)
1.06(40)
0.74(40)
0.81(37)
0.99(40)
1.71(36)
0.85(17)
1.07(62)
flux
(b)
0.46
0.73
0.07
0.25
0.26
0.23
0.26
0.23
0.25
0.23
0.23
0.38
2.65
0.18
0.87
6.99
11.88
11.43
10.80
10.34
10.77
11.59
8.67
11.94
0.56
0.57
0.55
0.57
2.56
0.28
0.05
0.04
0.50
0.12
0.51
0.10
0.06
1.6
1.2
1.1
0.22
0.79
0.21
0.12
0.55
0.59
0.61
0.59
0.59
0.19
1.14
(a) 1: single power law, 2: double power law, 3: broken power law, 4: logarithmic parabola model
(b) X-ray flux in the 2–10 keV band in units of 10−11 erg cm−2 s−1
Blazar Astrophysics with BeppoSAX and Other Observatories
1in11
Fig. 2. a- Spectral Energy Distribution of the BL Lacs 1ES 0033+595, PKS 0048−097, 1ES
0120+340 and RGB J0136+39
12
Giommi et al.
Fig. 2. b- Spectral Energy Distribution of the BL Lacs 1ES 0145+138, MS 0158.5+0019, 3C
66A and AO 0235+164
Blazar Astrophysics with BeppoSAX and Other Observatories
1in13
Fig. 2. c- Spectral Energy Distribution of the BL Lacs MS 0317.0+1834, 1H 0323+022, 1ES
0347−121 and 1H 0414+009
14
Giommi et al.
Fig. 2. d- Spectral Energy Distribution of the BL Lacs 1ES 0502+675, 1H 0506−039, PKS
0521−365 and PKS 0537−441
Blazar Astrophysics with BeppoSAX and Other Observatories
1in15
Fig. 2. e- Spectral Energy Distribution of the BL Lacs PKS 0548−322, S5 0716+714, MS
0737.9+7441 and OJ 287
16
Giommi et al.
Fig. 2. f- Spectral Energy Distribution of the BL Lacs B2 0912+29, 1ES 0927+500, 1ES
1028+511 and RGB J1037+57
Blazar Astrophysics with BeppoSAX and Other Observatories
1in17
Fig. 2. g- Spectral Energy Distribution of the BL Lacs RGB J1058+56, 1H 1100−230, Mkn
421 and 1RXS J111706.3+20141
18
Giommi et al.
Fig. 2. h- Spectral Energy Distribution of the BL Lacs EXO 1118.0+4228, Mkn 180, PKS
1144−379 and 1RXS J121158.1+22423
Blazar Astrophysics with BeppoSAX and Other Observatories
1in19
Fig. 2. i- Spectral Energy Distribution of the BL Lacs EXO 1215.3+3022, 1H 1219+301, ON
231 and 1RXS J123511.1−14033
20
Giommi et al.
Fig. 2. j- Spectral Energy Distribution of the BL Lacs 1ES 1255+244, MS 1312.1−4221,
EXO 1415.6+2557 and PG 1418+546
Blazar Astrophysics with BeppoSAX and Other Observatories
1in21
Fig. 2. k- Spectral Energy Distribution of the BL Lacs 1H 1430+423, MS 1458.8+2249, 1H
1515+660 and PKS 1519−273
22
Giommi et al.
Fig. 2. l- Spectral Energy Distribution of the BL Lacs 1ES 1533+535, 1ES 1544+820, 1ES
1553+113 and Mkn 501
Blazar Astrophysics with BeppoSAX and Other Observatories
1in23
Fig. 2. m- Spectral Energy Distribution of the BL Lacs 1ES 1741+196, S5 1803+784, 3C 371
and S4 1823+568
24
Giommi et al.
Fig. 2. n- Spectral Energy Distribution of the BL Lacs 1ES 1959+650, PKS 2005−489, PKS
2155−304 and BL Lacertae
Blazar Astrophysics with BeppoSAX and Other Observatories
1in25
Fig. 2. o- Spectral Energy Distribution of the BL Lacs 1ES 2344+514 and 1H 2354−315
26
Giommi et al.
Fig. 3. a- Spectral Energy Distribution of the FSRQs PKS 0208−512, NRAO 140, PKS 0523−33
and OG 147
Blazar Astrophysics with BeppoSAX and Other Observatories
1in27
Fig. 3. b- Spectral Energy Distribution of the FSRQs WGA J0546.6−6415, 1Jy 0836+710,
RGB J0909+039 and 1Jy 1127−145
28
Giommi et al.
Fig. 3. c- Spectral Energy Distribution of the FSRQs 3C 273, 3C 279, B3 1428+422 and GB
1508+5714
Blazar Astrophysics with BeppoSAX and Other Observatories
1in29
Fig. 3. d- Spectral Energy Distribution of the FSRQs PKS 1510−08, 1ES 1627+402, 3C 345
and V 396 HER
30
Giommi et al.
Fig. 3. e- Spectral Energy Distribution of the FSRQs 4C 62.29, S5 2116+81, PKS 2126−15 and
OX 57
Blazar Astrophysics with BeppoSAX and Other Observatories
1in31
Fig. 3. f- Spectral Energy Distribution of the FSRQs 1Jy 2149−306, PKS 2223+21, 3C 446
and CTA 102
32
Giommi et al.
Fig. 3. g- Spectral Energy Distribution of the FSRQs PKS 2243−123 and 3C 454.3
Blazar Astrophysics with BeppoSAX and Other Observatories
1in33
4. Summary and Conclusion
We have presented the X-ray spectrum and the Spectral Energy Distribution of a large
sample of blazars observed by BeppoSAX with the aim of providing a single homogeneous
reference for this type of BeppoSAX data. The collection of the results presented here
together with all the data is also available as part of the BeppoSAX archive at the ASI
Science Data Center (ASDC) at the following address
http://www.asdc.asi.it/blazars/
Given the heterogeneous nature of the sample we have not attempted to perform any
deep statistical studies. In the following we summarize our work and give some remarks
about possible interpretations. More detailed statistical studies or deeper interpretations
are reported elsewhere or will be the subject of future publications.
4. Spectral
energy distributions
a
Fig.
(PG 1418+546,
νpeak ≈ 0.4 eV ≈ 8 × 1013Hz) for which the X-ray emission is dominated by the flat inverse
Compton radiation and of an Intermediate BL Lac (ON 231, νpeak ≈ 1 eV ≈ 2 × 1014Hz) where
the simultaneous optical and BeppoSAX observations (Tagliaferri et al. 2000) clearly show that
the transition between the synchrotron and inverse Compton emission occurs in the soft X-ray
band.
typical LBL object
of
Fig. 5. Spectral energy distributions of HBLs where the X-ray emission is completely dominated
by synchrotron radiation. In the case of PKS 2155−304 νpeak is at ≈ 50 eV ≈ ×1016 Hz while
for the extreme HBL 1H 1430+423 νpeak is above 10 keV.
34
Giommi et al.
Fig. 6. The distribution of the synchrotron peak frequencies in the FSRQ and in the BL Lac
subsamples. While the νpeak values for FSRQs strongly cluster around 1014 Hz, a much wider
distribution is present in the BL Lacs subsample. In the latter case the distribution is strongly
biased towards high νpeak values by the BeppoSAX time allocation process which favoured
X-ray bright (and therefore high νpeak ) objects that could be detected by all NFI instruments.
The SED of the 84 blazars considered in this work confirms with large statistics
the widely accepted scenario where blazar emission is smooth across several decades
of energy and is characterized (in a νf (ν) vs ν representation) by two broad peaks
which are usually interpreted as being due to synchrotron emission followed by inverse
Compton radiation.
A wide range of X-ray spectral indices has been observed, ranging from very flat
values in Compton dominated sources like PG 1418+546 to very steep spectral slopes
in objects where the tail of the synchrotron emission just reaches the X-ray band. A
sharp transition from a very steep soft X-ray component to a much flatter hard X-ray
spectrum, marking the transition between the synchrotron and Compton emission, has
been clearly detected in a number of objects. Examples are given in Figures 4 and
5 where, for four representative objects, we plot the observed SED together with the
expected distribution from a SSC model peaking at appropriate νpeak values.
When viewed only from the narrow soft X-ray band, as was done in the past, these
SEDs clearly show that the local X-ray spectral index must be correlated to the X-ray
to radio flux ratio (fx/fr) as was first found by Padovani & Giommi (1996) and by
Lamer, Brunner & Staubert (1996) in large samples of BL Lacs observed with ROSAT.
Blazar Astrophysics with BeppoSAX and Other Observatories
1in35
The position of the synchrotron peak, estimated comparing the SEDs to SSC models
such as those shown in Figures 4 and 5, spans at least six orders of magnitudes ranging
from ≈ 0.1 eV in e.g. PKS 0048−097 or S5 2116+81 to 10–100 keV in some extreme
HBL BL Lacs like Mkn 501 and 1H 1430+423.
Very strong intensity and spectral variability can occur near the synchrotron (and
inverse Compton) peak. The position of this peak can move to higher energy by up to
two orders of magnitude (or perhaps more) during flares.
It is not clear what is the maximum νpeak that can be reached and whether Ultra
High energy synchrotron peaked BL Lacs (UHBL) exist. A few potential UHBL sources
may be present in the Sedentary survey (Giommi, Menna & Padovani 1999, Perri et al.
2002), which by definition only includes extreme HBL objects, especially those few that
are located within the error circle of unidentified EGRET sources. If these candidates
turn out to be the real counterpart of the EGRET gamma-ray sources their νpeak would
be so high that their synchrotron radiation would reach the gamma-ray band. One such
object, 1RXS J123511.1−14033 (see Figure 2i), was observed by BeppoSAX on three
occasions but always with short integration times giving inconclusive results (Giommi
et al. 2002, in preparation).
The observed distributions of νpeak values (rest frame), obtained by fitting SSC mod-
els to the multi-frequency data shown in Figure 2a-2o and 3a-3g, are plotted in Figure
6 for the FSRQ (top panel) and the BL Lac (bottom panel) subsamples. The two
distributions are certainly affected by selection effects, including that induced by the
BeppoSAX time allocation process which, by necessity, favoured high νpeak /X-ray strong
sources which could be detected by the high energy instruments. Although this bias is
clearly present in the BL Lac subsample where a large fraction of the sources are X-ray
bright HBL objects, in the case of FSRQs most of the objects have low νpeak . This
is because FSRQs are in general more luminous than BL Lacs and especially because
FSRQs with νpeak > 1016 Hz are very rare. To date the only FSRQ (RGB J1629+4008
= 1ES 1627+402, see Figure 3d) whose synchrotron emission reaches the X-ray band
was found by Padovani et al. 2002b.
A logarithmic parabola model, which can describe the spectral curvature of blazars
in a very wide energy band with only three parameters (see Landau et al. 1986), fits
better than other models (e.g. broken power law) the spectrum of HBL objects whose
X-ray emission is still due to synchrotron radiation. The average amount of spectral
curvature, as measured by the β parameter in the log parabola model of paragraph 3.1
is −0.38 +/− 0.1, a value somewhat steeper (possibly because of the energy dependant
synchrotron cooling), but not too different, than the amount of curvature found by
Landau et al. 1986 (−0.22 to −0.09 ) in a sample of BL Lacs whose synchrotron power
peaks at infra-red, optical frequencies. This similarity points to an intrinsically similar
curvature in the spectrum of the emitting particles. The smoothly changing slope could
be the spectral signature of a statistical acceleration mechanism where the acceleration
process becomes less and less efficient as the particle's energy increases (Massaro 2002).
In this scenario the widely different synchrotron νpeak energies in LBL and HBL objects
would be the result of the inefficiency in the acceleration process that sets off at different
energies.
Acknowledgements
This research has made use of data retrieved from the ASI/ASDC-BeppoSAX public
archive, the NASA/IPAC Extragalactic Database (NED), the NRAO VLA Sky Sur-
vey (NVSS), the Guide Star Catalog II (GSC2) and the Two Micron All Sky Survey
(2MASS).
36
References
Giommi et al.
Beckmann V., Wolter A. et al.: 2002, Astron. Astrophys. 383, 410
Ballo L., Maraschi L. et al.: 2002, Astrophys. J. 567, 50
Boella G. et al.: 1997a, Astron. Astrophys. Suppl. Ser. 122, 229
Boella G. et al.: 1997b, Astron. Astrophys. Suppl. Ser. 122, 327
Blandford R.D., Rees M.J.: 1978, in Pittsburgh Conf. on BL Lac objects, p. 328
Cardelli J.A., Clayton G.C., Mathis J.S.: 1989, Astrophys. J. 345, 245
Chiapetti L., Maraschi L. et al.: 1999, Astrophys. J. 521, 552
Costantini E., Comastri A. et al.: 1999, The Messenger 70, 265
Costamante L., Ghisellini G. et al.: 2001, Astron. Astrophys. 371, 512
Cusumano G., Mineo T. et al.: 2001, Astron. Astrophys. 375, 397
Dickey J.M., Lockman F.J.: 1990, Ann. Rew. Astron. Astrophys. 28, 215
Elvis M., Fiore F. et al.: 2000, Astrophys. J. 543, 545
Fabian A.C., Celotti A. et al.: 2001a, Mon. Not. R. Astr. Soc. 323, 373
Fabian A.C., Celotti A. et al.: 2001b, Mon. Not. R. Astr. Soc. 324, 628
Frontera F. et al.: 1997, Astron. Astrophys. Suppl. Ser. 122, 357
Fiore F., Guainazzi M., Grandi P.: 1999, Cookbook for NFI spectral analysis
Fossati G. et al.: 2000a, Astrophys. J. 541, 153
Fossati G. et al.: 2000b, Astrophys. J. 541, 166
Ghisellini G., Costamante L. et al.: 1999, Astron. Astrophys. 348, 63
Giommi P. et al.: 1990, Astrophys. J. 356, 432
Giommi P., Fiore F.: 1997, Data Analysis in Astronomy IV, p. 93
Giommi P., Fiore F. et al.: 1998, Astron. Astrophys. Lett. 333, L5
Giommi P., Massaro E. et al.: 1999, Astron. Astrophys. 351, 59
Giommi P., Menna M.T., Padovani P.: 1999, Mon. Not. R. Astr. Soc. 310, 465
Giommi P., Padovani P., Perlman E.: 2000 Mon. Not. R. Astr. Soc. 317, 743
Jager R. et al.: 1997, Astron. Astrophys. Suppl. Ser. 125, 557
Lamer G., Brunner H., Staubert R.: 1996, Astron. Astrophys. 311, 384
Landau R. et al.: 1986, Astrophys. J. 308, 78
Malizia A., Capalbi M. et al.: 2000, Mon. Not. R. Astr. Soc. 312, 123
Manzo G. et al.: 1997, Astron. Astrophys. Suppl. Ser. 122, 341
Massaro E., Cusumano G., Litterio M., Mineo, T.: 2000, Astron. Astrophys. 361, 695
Massaro E., Giommi P. et al.: 2002, Astron. Astrophys. submitted
Massaro E.: 2002, these proceedings
Morrison R., McCammon D.: 1983, Astrophys. J. 270, 119
Mineo T., Fiore F. et al.: 2000, Astron. Astrophys. 359, 471
Padovani P., Giommi P.: 1996, Mon. Not. R. Astr. Soc. 279, 526
Padovani P. et al.: 2001, Mon. Not. R. Astr. Soc. 328, 931
Padovani P. et al.: 2002a, Mon. Not. R. Astr. Soc. submitted
Padovani P. et al.: 2002b, Astrophys. J. submitted
Perri M., Giommi P., Piranomonte S., Padovani P.: 2002, these proceedings
Pian E. et al.: 1998, Astrophys. J. Lett. 492, L17
Pian E. et al.: 2002, Astron. Astrophys. 392, 407
Parmar A.N. et al.: 1997, Astron. Astrophys. Suppl. Ser. 122, 309
Ravasio M., Tagliaferri G. et al.: 2002, Astron. Astrophys. 383, 763
Siebert J., Brinkmann W. et al. : 1999, Astron. Nach. 320, 315
Tagliaferri G., Ghisellini G., Giommi P. et al.: 2000 Astron. Astrophys. 354, 431
Tavecchio F., Maraschi L., et al.: 2000, Astrophys. J. 543, 535
Tavecchio F. et al.: 2002, Astrophys. J. 575, 137
Tanihata C., et al.: 2001 Astrophys. J. 563, 569
Urry M.C., Padovani P.: 1995, Publ. Astr. Soc. Pacific 107, 803
Wolter A., Comastri A. et al.: 1998, Astron. Astrophys. 335, 899
Zhang Y.H., Celotti A. et al.: 1999, Astrophys. J. 527, 719
|
astro-ph/9912078 | 1 | 9912 | 1999-12-03T22:12:23 | A speedy pixon image reconstruction algorithm | [
"astro-ph"
] | A speedy pixon algorithm for image reconstruction is described. Two applications of the method to simulated astronomical data sets are also reported. In one case, galaxy clusters are extracted from multiwavelength microwave sky maps using the spectral dependence of the Sunyaev-Zel'dovich effect to distinguish them from the microwave background fluctuations and the instrumental noise. The second example involves the recovery of a sharply peaked emission profile, such as might be produced by a galaxy cluster observed in X-rays. These simulations show the ability of the technique both to detect sources in low signal-to-noise data and to deconvolve a telescope beam in order to recover the internal structure of a source. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 4 September 2017
(MN LATEX style file v1.4)
A speedy pixon image reconstruction algorithm
Vincent Eke1,2
1Institute of Astronomy, Madingley Road, Cambridge. CB3 0HA, UK
2Steward Observatory, 933 N Cherry Ave, Tucson. AZ 85721, USA
4 September 2017
ABSTRACT
A speedy pixon algorithm for image reconstruction is described. Two applications of
the method to simulated astronomical data sets are also reported. In one case, galaxy
clusters are extracted from multiwavelength microwave sky maps using the spectral
dependence of the Sunyaev-Zel'dovich effect to distinguish them from the microwave
background fluctuations and the instrumental noise. The second example involves the
recovery of a sharply peaked emission profile, such as might be produced by a galaxy
cluster observed in X-rays. These simulations show the ability of the technique both
to detect sources in low signal-to-noise data and to deconvolve a telescope beam in
order to recover the internal structure of a source.
Key words: methods: data analysis -- techniques: image processing -- galaxies: clus-
ters -- cosmic microwave background -- X-rays: general
1
INTRODUCTION
The process of measuring an image can, for many applica-
tions in astronomy, be written as
D(x) = (T ∗ B)(x) + N (x).
x represents a pixel position in a two-dimensional array, or
image, D is the observed data, T is the true image, B is
the point-spread function (psf) of the measuring instrument,
assumed to be invariant across the image, with N being the
associated statistical noise and ∗ represents the convolution
operator such that
(f ∗ g)(x) = Zn
f (y)g(x − y)dy.
(1.1)
(1.2)
n is defined to be the number of pixels in the data image. The
task of an image reconstruction algorithm is to infer the true
image given the data, knowledge of the psf and the statistical
properties of the noise. For the simple case when there is no
noise then, assuming that the inverse of the beam exists,
that there is no structure on sub-pixel scales and ignoring
complications introduced by edge effects, the inferred truth,
T , can be obtained using the convolution theorem and will
exactly equal T . Denoting the Fourier transform of f (x) by
f (k), T can be found by inverse transforming
T (k) = D(k)/ B(k).
(1.3)
The n pieces of information in the data allow a perfect re-
construction of the n pixel values in the truth. With noise
switched on, the number of degrees of freedom in the solu-
tion increases by n, while the number of constraints remains
c(cid:13) 0000 RAS
fixed, yielding an ill-posed problem. Thus, the job of the re-
construction algorithm becomes to decide which of the pos-
sible inferred truths is the best, whatever that may mean.
An extension of the simple case described above involves
filtering the data to remove the noise before inverse trans-
forming equation (1.3) to yield T . One particular, widely
used example of this is the Wiener filter (Wiener 1949; Ry-
bicki & Press 1992; Lahav et al. 1994; Bunn et al. 1994;
Fisher et al. 1995). This procedure is non-iterative (it iter-
ates to T = 0) and the final inferred truth is completely
determined once guesses for the power spectra of the true
signal and the noise components have been made. If the as-
sumed power spectra are correct then the resulting filter will
minimise over the whole image the variance between the re-
constructed and true signals. It can also be shown that the
Wiener filter yields the maximally probable inferred truth
if the true and noise values are normally distributed (Ry-
bicki & Press 1992). For situations where these distribu-
tions are not Gaussian, the optimal filter differs from the
Wiener filter. While the Wiener filter method involves only
a few transforms, and is thus rapid, it would in general be
desirable to break the degeneracy in solution space with a
technique that both does not require any assumption about
the nature of T and produces optimal images for any true
signal distribution.
Another type of transform is the wavelet transform (e.g.
Slezak, Bijaoui & Mars 1990) where the inferred truth is
described using a set of basis functions designed to extract
information on a variety of scales. Once again, this is a linear
method that involves applying a transformation to the data
2
V. Eke
in order to extract information about particular scales of
interest.
An alternative approach to image reconstruction in-
volves quantifying the acceptability of a particular inferred
truth, then iterating the inferred truth until it becomes max-
imally acceptable. One reasonable demand to make of T is
that the resulting distribution of residuals, defined as
R(x) = D(x) − ( T ∗ B)(x),
is statistically indistinguishable from that of the anticipated
noise N , i.e. requiring a 'good fit' to the data. The partic-
ular statistic used to describe the extent of the misfit will
depend on the nature of N . For example, χ2 is a frequently
employed statistic when noise values are drawn from a nor-
mal distribution.
(1.4)
Having chosen a misfit statistic, approaches to reducing
further the acceptable solution space are more varied. A
common and simple route is to parametrise T and use the
data to fit a small number of parameters. This is very quick
and effective, provided that the prejudice contained in the
parametrisation is appropriate. For complicated images, a
more sophisticated procedure is desirable.
To understand better where pixons fit into the story,
consider the following conditional probability equation with
M representing all aspects of the model used to transform
T to D:
.
(1.5)
(1.6)
p(D)
p(D T , M )p( TM )p(M )
p( T , MD) =
This is merely the probability of having a particular com-
bination of data, model and inferred truth, divided by the
probability of obtaining a particular data set. Since D is
measured before it is used to infer T , p(D) is constant. In
addition, the desire to avoid introducing prejudice requires
that p(M ) is constant over all models. This leaves
p( T , MD) ∝ p(D T , M )p( TM ).
On the left hand side of this proportionality is the quan-
tity that one would like to maximise in the reconstruction,
namely the probability of a combination of inferred truth
and model given the data. The first term on the right hand
side is readily identified as the 'likelihood', and the second
term is commonly called the 'image prior'. From a Bayesian
viewpoint, it is reasonable to associate a probability to the
plausibility of obtaining a particular inferred truth once a
model is specified, despite the non-repeatable nature of the
image prior. This provides additional leverage in the quest to
reduce the acceptable solution space. Note that even if the
model is held fixed and one seeks to maximise p( TM, D)
the above equation looks very similar because in that case
p(MD) = 1 and p( T , MD) = p( TM, D)p(MD).
In order to quantify the image prior, consider the situ-
ation of distributing Z indistinguishable photons randomly
among q buckets. Denoting the number of photons in bucket
i by zi, the probability of a particular distribution is
.
(1.7)
p({zi}) =
Z!
i=1 zi!
qZ Qq
The most probable choice of {zi} can readily be shown to be
the one that has the same number of photons in each bucket.
This probability also increases when the number of buckets
is decreased; an Occam's razoresque tendency to favour sim-
ple descriptions. Referring back to the image prior, p( TM )
will be increased by having the photons distributed evenly
through the model buckets, and by reducing the number of
buckets.
This approach to image reconstruction was pioneered
by Skilling (1989) and Gull (1989). By analogy with sta-
tistical thermodynamics, they related the maximisation of
the image prior to a maximisation of the entropy of the re-
constructed truth. In the absence of additional information
which could shift the prior away from the default flatness,
uniform T s are preferred over more complicated images that
require a less probable distribution of the indistinguishable
photons in the image pixels.
These maximum entropy methods (MEM) are conven-
tionally applied in the pixel grid in which the data are mea-
sured. However, if the real truth deviates from flatness then,
to the extent that the image prior is important relative to the
likelihood term in equation (1.6), this procedure will bias T
away from T . This suggests that the distribution of buckets
into which the photons are placed should be set according
to the measured distribution D if the prior probability is to
be truly maximised. Furthermore, one implication from the
above discussion is that the number of buckets should be
minimised in order to create the simplest, and consequently
most plausible, description for T , rather than using n pa-
rameters because this is how many pixels there are in the
data image. Considerations such as these led Pina & Puetter
(1993, hereafter PP93) to introduce the pixon concept. Un-
like the uniformly arranged pixels, pixons are able to adapt
to the measured D in order that the information content is
flat across the inferred truth when described in the pixon
basis. This adaptive 'grid' essentially uses a higher density
of pixons to describe the inferred truth in regions where
more signal exists, and only a few very large independent
pixons for the background, or low signal-to-noise parts of
an image. Relative to MEM, the pixon method allows the
probability in equation (1.6) to maximise itself with respect
to one aspect of the model which conventional MEM keep
fixed. The task of maximising p( T , MD) boils down to find-
ing the inferred truth containing the fewest pixons, which
nevertheless provides an acceptable fit to the data. Opera-
tionally, the main difference between MEM and the pixon
method is that MEM explicitly quantifies the image prior
and does a single maximisation of p( TM, D). In the pixon
case this is split into successive likelihood and image prior
maximisations. Both MEM and the pixon method require
an assumption to be made concerning the noise distribution
in order that the goodness-of-fit can be quantified.
In summary, the pixon method is an iterative image
reconstruction technique that produces inferred truths that
are smooth on a locally defined pixon scale, and the itera-
tions have a well-determined finishing point, namely when
the pixon distribution has converged to the simplest state
that yields an acceptable fit to the data. This approach to
image reconstruction has been applied to a variety of astro-
nomical data sets (e.g. Smith et al. 1995; Metcalf et al. 1996;
Dixon et al. 1996, 1997; Knodlseder et al. 1996). A detailed
discussion of the theoretical basis of the pixon, as well as a
list of applications of the method can be found in the paper
by Puetter (1996).
In the original pixon implementation described by
PP93, the time taken to perform a reconstruction scaled
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
with the square of the total number of pixels in the im-
age. For typical astronomical images this meant reconstruc-
tions would be impractically slow. While the discussion so
far shows in principle the advantages offered by the pixon
approach, the implementation of the idea remains to be spec-
ified. The main purpose of this paper is to describe a speedy
pixon algorithm that is limited by the calculation of fast
Fourier transforms, thus reducing the scaling of the run time
to nlogn and allowing 2562 pixel images to be reconstructed
in a few minutes on a workstation. In Section 3 the method is
applied to two simulated astronomical data sets. The results
are compared with a simple maximum likelihood method
and the robustness of the reconstructions is tested quan-
titatively using Monte Carlo simulations. Puetter & Yahil
(1999) have reported the existence of accelerated and quick
pixon methods.
2 METHOD
2.1 Outline
The speedy pixon algorithm does not differ greatly from
that originally proposed by PP93. Namely, the maximisa-
tion of the probability in equation (1.6) is performed in an
iterative fashion with each step consisting of a change in the
number of pixons being used to describe the inferred truth
followed by a conjugate gradient maximisation of the like-
lihood (or equivalently minimisation of the misfit statistic)
for the fixed pixon distribution. These iterations are stopped
when the inferred truth is described using the fewest pixons
that allow an adequate fit to the data. The details of the
pixon implementation are given in the following section, be-
fore considering how they impact upon the minimisation of
the misfit statistic. A flow diagram outlining the entire pro-
cedure is shown in Figure 1.
2.2 Specifics of the pixon implementation
For the examples described in Section 3 the inferred truth is
reconstructed in the same grid as the observed data and thus
contains n values, one for each pixel. Given that the pixons
are to be less numerous than the pixels, these n evaluations
of the inferred truth cannot be independent. In practice a
pseudoimage H is defined in the pixel grid and T is evaluated
by correlating the pseudoimage values over a region with a
size determined by the local information content. This can
be written as
T (x) = Zn
P (x, y)H(y)dy
(2.1)
where P (x, y) is a weight function, the pixon shape, defin-
ing how much signal the inferred truth at pixel x gathers
from the pseudoimage at y. From this equation it can be
seen that T (x) merely represents the result of convolving
the pseudoimage with the pixon shape appropriate to pixel
x. Thus, provided the number of distinct pixon shapes is
finite and independent of n, it is already apparent why the
computational time for this method is going to scale like
nlogn rather than the direct summation n2 of the original
method.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Speedy pixon image reconstruction
3
The choice of pixon shapes and sizes will place con-
straints on the types of truths that could possibly be recov-
ered. Thus, using the vocabulary adopted by Puetter (1996),
it is important that the richness in the pixon language is suf-
ficient to enable a wide variety of truths to be reconstructed.
After all, it should be the data which drives the reconstruc-
tion algorithm to an inferred truth, rather than the algo-
rithm knowing beforehand what it is going to see! For the
reconstructions presented in Section 3, a set of npixon cir-
cularly symmetric 2-dimensional Gaussians was used with
widths, {δl : l = 1, npixon}; δl+1 > δl, truncated at r = 3δ.
This pixon shape was found to give better results than ei-
ther an inverted paraboloid or a top hat for the examples
considered.
Constructing T in this way it was found that, for the ex-
amples described in Section 3.2 where there are large ranges
in signal strength across the image, the vast majority of
pseudoimage pixels were being set to zero during the recon-
struction. Consequently the T values in these pixels were
determined by the distance to the few pseudoimage pixels
that contained any signal and the appropriate pixon weight.
To provide a little more flexibility to the pseudoimage in
such a situation, rather than relying on the richness in the
pixon set to enable a variety of images to be inferred from
a near delta function pseudoimage, the normalisation of the
lth pixon was chosen such that
Zn
δl (cid:17)ψ
Pl(x, y)dy = (cid:16) δ1
.
(2.2)
With ψ > 0, this means that in pixels having large pixon
widths, the inferred truth will be suppressed relative to that
calculated when all pixons are normalised to the same value.
The pseudoimage values in the low signal regions, where the
larger pixons are being used, are thus encouraged to increase
relative to those in the high signal parts of the image. An
unpleasant side effect of this is that the T values in high
signal pixels can become unduly influenced by the boosted
regions of the pseudoimage, so in calculating the signal, the
following weights were also included:
W (δ(x), δ(y)) = (cid:26) 1
(δ(x)/δ(y))ψ
if δ(x) ≥ δ(y)
if δ(x) < δ(y)
(2.3)
such that
T (x) = Zn
P (x, y)W (δ(x), δ(y))H(y)dy.
(2.4)
This down-weights the contribution to T gathered from the
pseudoimage pixels with larger δs. Values of ψ between 0
and 1 were used for the reconstructions in Section 3. The
introduction of this extra weight means that the calculation
of T no longer involves just a single convolution. However,
by splitting the integral into npixon separate convolutions of
Pl with a W (δl, δ(y))-weighted H, the nlogn scaling can be
preserved.
The procedure for defining the distribution of pixon
widths is based on the desire to have the same amount
of information in each pixon. Information content can be
parametrised through a pixon signal-to-noise ratio (SNR).
This is defined as
4
V. Eke
Double the pixon SNR
Read in data
Choose minimum pixon widths
and minimise misfit
Select a large pixon SNR
Calculate pixon width distribution
Minimise misfit
yes
Acceptable fit?
no
Interval bisect to give new pixon SNR
and recalculate the pixon widths
Minimise misfit
Acceptable fit?
no
yes
Reload information from the
last oversmoothed attempt
Stop
yes
Has pixon SNR converged?
no
Figure 1. Flow diagram showing the structure of the algorithm. There is an initial maximum likelihood type of fit, with all pixon widths
taking the minimum available value. The resulting pseudoimage is used to find a pixon SNR that is too large to enable an acceptable fit
to be found. Then the largest pixon SNR giving rise to an acceptable residual distribution is found by interval bisection.
S(δ(x)) =
δ1 (cid:19)
P (x, y)σ2(y)dy (cid:18) δ(x)
δ(x)Rn P (x, y) T (y)dy
qRn
ψ
2
.
σ(y) is the anticipated amplitude of the noise in pixel y.
The final term just represents the removal of the non-
normalisation of the pixons, and the first δ(x) is there be-
cause the mean SNR within the pixon is being calculated.
For a specified pixon SNR, the pixon width distribution is
chosen such that all pixels have the minimum available δ
that provides at least the required SNR.
For the reconstructions described in Section 3.2, it was
found that relaxing the need for the inferred truth to have
a constant SNR in each pixon led to superior results. With
a flat inferred truth in the pixon basis, the higher signal
regions of the data image received larger, and thus less likely,
reduced residuals. In order to decrease this discrepancy, the
required pixon SNR was decreased in the high signal pixels.
(2.5)
The pixon SNR appropriate for pixel x was defined to be
fSNR(x) times the global value, where
fSNR(x) = max(υ,pmax(0., 1 − (1 − υ2)r(x))),
(2.6)
with r(x) being the convolution of R(x)/σ(x) with the cor-
rectly normalised pixon appropriate to each pixel, and υ a
constant chosen to lie in the range [0, 1]. While this is a very
ad hoc addition to the method, it encapsulates the desire
to provide more freedom in the badly fitting regions of the
reconstructed image, rather than keeping rigorously to the
maximum entropy flat T solution.
The remaining part of the pixon story, is to describe
the fashion in which the pixon SNR is iterated. To define
the first T , in order that the SNR can be computed via
equation (2.5), a maximum likelihood type fit is performed
using δ(x) = δ1 ∀x and starting from a flat T with the mean
value of D. An initial pixon SNR is chosen such that the
resulting pixon distribution has too few degrees of freedom
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
to allow a good fit to be obtained. Interval bisection is then
used to find the maximum acceptable pixon SNR. At each
stage, the last over-smoothed (i.e. badly fitting) T is used to
infer the new pixon distribution. This decreases the chance
of introducing spurious structures into the reconstruction.
In practice, once the pixon SNR converges to about 20 per
cent, T is insensitive to further refinement and the procedure
is stopped.
2.3 Specifics of the likelihood calculation
The other half of the reconstruction method involves the
likelihood term, and the procedure used to maximise this
for a fixed pixon distribution. A likely T is one with a small
misfit statistic, ie one that 'fits the data well'. Gaussian and
Poisson distributed noise are relevant for the examples in
Sections 3.1 and 3.2, so reconstructions were attempted us-
ing the χ2 and
χ2
γ =
n
Xi=1
(R(i) + min(D(i), 1))2
D(i) + 1
(2.7)
(Mighell 1999) misfit statistics respectively. However, supe-
rior results were obtained by employing the ER statistic of
Pina & Puetter (1992, PP92). Rather than just taking into
account the amplitudes of the residuals, this measures their
spatial autocorrelation function. In more detail, ER is de-
fined as
ER =
1
n Xz
AR(z)2,
(2.8)
where z represents a 2-dimensional lag in pixel space and
AR is the autocorrelation of the reduced residuals
AR(z) = (R/σ ⊗ R/σ)(z)
= Xn
(R/σ)(y + z)(R/σ)(y).
(2.9)
As PP92 showed, the expected value of ER is equal to the
number of lags included in the summation in equation (2.8),
and the extent over which lag terms are useful is determined
by the size of the instrumental psf. For the applications de-
scribed below, an acceptable fit is defined to be one that has
ER < 3 + the number of lag terms being considered. This
is true ∼ 90 per cent of the time for normally distributed
noise or Poisson distributed noise when the mean signal is
∼> 1/√n.
To minimise the misfit statistic, the Polak-Ribiere con-
jugate gradient algorithm in Numerical Recipes (Press et al.
1992) was used. Some alterations were made to tailor the
general purpose routine to this specific task, as detailed in
Appendix A. In order to have non-negative inferred truths,
PP92 suggested using transformed pseudoimage values, Ht,
in the minimisation where
Speedy pixon image reconstruction
5
The calculation of the gradient of the misfit statistic
with respect to the transformed pseudoimage values is a bit
messy. After all, changing Ht, or effectively H, alters the
inferred truth in surrounding pixels. This is then convolved
with the instrumental psf before the residuals and hence the
misfit are calculated. Appendix B contains the results of this
calculation, but the important point is that it can be split
up into correlations and convolutions, thus maintaining the
nlogn scaling of the algorithm.
3 APPLICATIONS
The speedy pixon algorithm was applied to two simulated
astronomical data sets. In the first case, the challenge of
identifying galaxy clusters in Cosmic Microwave Background
(CMB) maps was considered for an instrument with speci-
fications like those of the Planck surveyor. The formalism
described above will be extended to deal with the mul-
tifrequency and multicomponent natures of the data and
truth respectively. In the second example some simulated
'β'-profiles convolved with a large psf, such as the ASCA X-
ray detector would measure when pointed at galaxy clusters,
are reconstructed.
3.1 Multiwavelength cluster detection in
simulated CMB data
The Planck surveyor satellite (Tauber, Pace & Volont´e 1994)
is expected to return maps of the sky in a number of mi-
crowave wavelength ranges. In addition to the intrinsic CMB
fluctuations, a number of interesting foregrounds will also
contribute to these maps. One such contribution will come
from the Sunyaev-Zel'dovich (SZ) effect produced when
CMB photons are inverse Compton scattered during their
passage through the ionised gas in galaxy clusters (Sunyaev
& Zel'dovich 1972). The distinctive spectral distortion cre-
ated by the net heating of CMB photons in the directions
of galaxy clusters should enable some thousands of clusters
to be detected by Planck (Hobson et al. 1998).
3.1.1 Additional formalism
The formalism described here will assume that the spectral
dependence of each of the components (i.e. CMB, SZ and
noise) is known and constant over the region being observed,
although in principle this could also be left as a part of the
model over which the probability is maximised. For each of
the nc components, the spectral template will be denoted
by ωc(j) : c = 1, nc; j = 1, nλ, where nλ is the number of
wavebands in the observed data. Thus the inferred truth in
waveband j can be written
H(x) = α(Ht(x) +pHt(x)2 + β),
with α = 0.5 and β = 1. This transformation was used for
the example in Section 3.1, but created a poor reconstruction
of the low-signal regions in the examples in Section 3.2. For
these cases, setting Ht = H and simply truncating negative
values yielded superior results without significantly affecting
the speed of the code.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
(2.10)
T (x, j) =
nc
Xc=1
ωc(j)Zn
Pc(x, y)W (δc(x), δc(y))Hc(y)dy. (3.1)
Each component has its own pixon distribution and pseu-
doimage denoted by the c subscript. The pseudoimage vari-
ables for the intrinsic CMB and thermal SZ components
are the thermodynamic ∆T /T and the Comptonisation pa-
rameter y respectively, both in units of 10−6. Note that
6
V. Eke
Table 1. Lists of (1) the central Planck observing frequencies employed here (GHz); (2) the Gaussian full width half maxima (arcmin); (3)
the 1σ Gaussian noise per 1.5 arcmin square pixel in mJy (for 14 months of observations); (4) the conversion factor relating thermodynamic
CMB ∆T /T , in units of 10−6, to the change in flux in mJy in a pixel; (5) the conversion factor relating the Comptonisation parameter
y, in units of 10−6, to the change in flux in mJy in a pixel; (6) the rms cluster T ∗ B per pixel (mJy); (7) the maximum amplitude of
cluster T ∗ B per pixel (mJy); (8) the rms intrinsic CMB T ∗ B per pixel (mJy); (9) the maximum amplitude of intrinsic CMB T ∗ B per
pixel (mJy).
(1)
Frequency
100
143
217
353
(2)
(3)
fwhm Noise
10.7
8.0
5.5
5.0
1.2
2.2
3.1
5.8
(4)
ωSZ(j)
-0.19
-0.20
-1.9×10−3
0.34
(5)
(6)
(7)
(8)
(9)
ωCMB(j)
rms TSZ ∗ B max TSZ ∗ B
rms TCMB ∗ B max TCMB ∗ B
0.124
0.197
0.249
0.155
0.20
0.23
2.3×10−3
0.43
5.1
7.8
0.11
20.6
4.6
7.6
9.9
6.2
17.4
29.1
38.5
24.1
∆T /T can take positive or negative values so the trans-
formed pseudoimage is chosen to equal the pseudoimage for
this component, whereas equation (2.10) is used to ensure
that HSZ(x)(≡ y) is non-negative in all pixels. For the re-
constructions presented below, the value of ψ, as defined in
equation (2.2), was set to zero such that all pixons were nor-
malised to unity. In practice, it may be beneficial to allow
ψ to be a function of component.
The multiwavelength data, and the fact that the CMB
component can take positive or negative values, require a
definition of pixon SNR differing from that in equation (2.5).
This is chosen to be
those lying outside the input data image had R(x) = 0 de-
fined in them, but were otherwise treated identically with
the rest of the reconstructed image pixels. This buffer re-
gion allows the pixon SNR to be sensibly defined, so that
the inferred truths have the same sensitivity across all of the
observed image while not directly influencing the calculation
of the misfit statistic.
The computation of the misfit statistic includes nλ dif-
ferent residual maps and the definition of an acceptable
value is modified accordingly. In addition, the misfit min-
imisation is performed over n × nc variables.
S(δc(x)) =
ψ
2
.
(3.2)
3.1.2 Data production
Pc(x, y) Tc(y, 1)dy
δ1,c (cid:19)
Pc(x, y)σ2(y, 1)dy (cid:18) δc(x)
δc(x)Rn
qRn
The 1 index with the T and σ2 terms represents that only
the first waveband is being used to define the signal-to-noise
ratio and δ1,c is the smallest pixon width for component c.
For the type of data simulated here, with the intrinsic CMB
component dominating the thermal SZ component, it is pos-
sible, if the required pixon SNR is the same for all compo-
nents, to find an acceptable fit to the data without the need
for a cluster component in the inferred truth. In order to al-
low a better reconstruction of the weaker component (i.e. a
reconstruction that includes some signal), it is necessary to
introduce factors gSNR(c) such that the actual pixon SNR
requested for pixons representing component c is gSNR(c)
times the default value. While these factors could be left
for the pixon algorithm to evaluate in a Bayesian fashion,
this would be rather time consuming. In practice the relative
strengths of the components were set to gSNR = 1 and 0.1
for the intrinsic CMB and thermal SZ components respec-
tively in order that clusters were found, without introducing
many spurious sources.
The υ parameter for determining the variation in pixon
SNR across the image for a given component was set to 1
for both the CMB and SZ components. This enforced a flat
pixon SNR across each of the component maps (fSNR = 1).
If some SNR variation was to be used, then an average over
wavebands of the reduced residuals convolved with the local
pixon shape would need to be calculated, rather than the
monochromatic version contained in equation (2.6).
To deal with the sharp edge in the observed data, the re-
constructed images were allowed to extend beyond the orig-
inal data pixel grid. A total of 5122 pixels were used and
A 4002 pixel, 10 degree square field of simulated CMB sky,
was created including both intrinsic CMB fluctuations and
thermal SZ distortions produced by clusters, The intrinsic
CMB map was a realisation of the standard Cold Dark Mat-
ter model using the power spectrum returned by CMBFAST
(Seljak & Zaldarriaga 1996). The thermal SZ map was pro-
duced by creating some templates from the hydrodynamical
galaxy cluster simulations of Eke, Navarro & Frenk (1998)
and then pasting these, suitably scaled, at random angu-
lar positions with mass and redshift distributions according
to the Press-Schechter formalism (Press & Schechter 1974).
The thermal SZ Comptonization parameter, y, and the ther-
modynamic ∆T /T of the intrinsic CMB fluctuations were
converted to fluxes per pixel in mJy⋆ in each of four wave-
bands by multiplying by ωc(j) as listed in columns 4 and 5 of
Table 1. These four combined truths were then 'observed' by
applying the relevant Gaussian beam and adding the pixel
noise appropriate to each wavelength (see columns 2 and 3
of Table 1).
The final four columns in Table 1 show the maximum
Tc∗B and rms Tc∗B per pixel produced by each of the two
components separately. These numbers demonstrate both
the high SNR with which Planck will observe the intrinsic
CMB fluctuations and the relatively low SNR of even the
brightest cluster in the field, after convolution with the in-
strumental psfs.
⋆ 1 mJy ≡ 10−29 W m−2 Hz−1
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Speedy pixon image reconstruction
7
Figure 2. The top left panel contains a histogram of the difference between the true and pixon-inferred intrinsic CMB component fluxes
at 100 GHz in each pixel. Flux units are mJy for all panels in this figure. The width of the best-fitting Gaussian is also given. In the
lower left panel, the solid line represents the average difference between the true and pixon-inferred intrinsic CMB component fluxes as a
function of the true pixel flux, and the dashed line traces the standard deviation of the reconstruction error. The two right hand panels
show the corresponding results for the ML reconstruction. In both cases, the Gaussian fits to the reconstruction errors are not shown
because they essentially lie on top of the histograms.
3.1.3 Results
The sets of pixon widths for the two components were cho-
sen to be 3δCMB(x) = 5, 7, 9 and 3δSZ(x) = 5, 9, 80. For the
intrinsic CMB, the signal-to-noise ratio is so good that the
pixons do not need to be very large before they prevent an
acceptable fit from being found. In the SZ clusters case, the
selection of a small and a very large pixon size allows the re-
construction of sharp features in a smooth background. The
intermediate-sized pixon helps the pixons bridge the gap be-
tween background and cluster as the pixon SNR is reduced
and the map becomes progressively less correlated. These SZ
pixon widths are selected to match the scales of the features
expected to be present for this component map. About 20
minutes of computer time were required to perform the mul-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
8
V. Eke
Figure 3. The true thermal SZ y map is shown in the top panel,
and the ML and pixon-inferred truths are contained in the second
and third panels respectively. Axes are labelled in pixels.
ticomponent and multiwavelength speedy pixon reconstruc-
tion of the 4002 pixel image (padded up to 5122 pixels). For
comparison a maximum likelihood (ML) reconstruction with
all pixons being set to be pixel-sized was also performed.
This took approximately 3 minutes to complete.
In Figure 2, the pixon and ML-inferred truths for the
intrinsic CMB flux at 100 GHz are compared with the true
values in each pixel. The top panels show the distributions
of reconstruction errors per pixel for the two methods, along
with the widths of the best-fitting Gaussians for these distri-
butions. Comparing the raw data (i.e. including SZ clusters,
convolution with the beam and noise) with the true intrinsic
CMB pixel values, leads to a best-fitting Gaussian width of
1.70 mJy, so it is apparent that both pixon and ML recon-
structions have cleaned the data to some extent, although
the narrowing of the error distribution is significantly better
for the pixon case. The lower panels show trends for both
the mean (solid lines) and standard deviation (dotted lines)
of the flux errors as a function of the true pixel flux. In the
ML case, the scatter in the flux errors is large and approxi-
mately independent of the true signal, whereas for the pixon
reconstruction the scatter is suppressed but increases when
the signal is strong and the pixon width being used becomes
smaller. Where the scatter in the error increases, the mean
difference between the pixon-inferred and true signals de-
creases. This shows that for pixels with absolute values of
intrinsic CMB 100 GHz flux exceeding ∼ 6 mJy, a smaller
pixon width has been selected and the fit has improved. The
choice of pixon widths is thus very important in determining
these results. In the ML case, the trend in the mean error
shows that the peak sizes are systematically underestimated.
Figure 3 is a greyscale comparison of the true (top
panel), ML-inferred (middle) and pixon-inferred (bottom)
cluster y maps. It is very apparent that the pixon recon-
struction has greatly suppressed the noise relative to the
ML effort. There are a few sources in the pixon reconstruc-
tion that do not correspond with single identifiable sources
in the actual truth. In regions where the density of small
clusters is particularly high, the pixon algorithm has a ten-
dency to place a single bright source to model the emission.
However, relative to the ML effort, the compression of the
reconstructed information is very clear. The pixon algorithm
has essentially already made the decision as to which of the
many ML sources are statistically justifiable. Reducing the
SZ pixon SNR relative to that of the intrinsic CMB using
the gSNR(SZ) parameter would allow the pixon algorithm to
detect clusters with smaller fluxes, albeit with an increased
risk of producing spurious sources. The mean Compton y
parameters per pixel in units of 10−6 are 0.80, 1.25 and 0.79
for the true, ML and pixon images respectively, so the pixon
algorithm does a good job of conserving the entire thermal
SZ flux, in contrast to the ML technique.
As an aside, the inclusion of an intrinsic CMB compo-
nent does not affect the cluster detection efficiency signif-
icantly. The important quantity is the signal-to-noise ratio
with which the clusters alone would be observed. Other fore-
grounds such as dust and Galactic free-free and synchrotron
emissions are unlikely to vary on small scales and should not
greatly affect the ability of the algorithm to detect clusters
(see e.g. Hobson et al. 1998).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Speedy pixon image reconstruction
9
Figure 4. Contour plots for the X-ray cluster example in Section 3.2 showing the central region of the high signal-to-noise true β profile
(top left), one noisy realisation of it (lower left), a pixon-inferred truth (top right) and the ML-inferred truth (lower right). The contour
levels are 0.5, 1, 3, 10, 30 (bold),70, 150 and 300 counts per pixel and the scales on the axes are numbers of pixels.
3.2 Simulated ASCA X-ray cluster data with
Poisson distributed noise
X-ray imaging of galaxy clusters using the ASCA satellite
involves convolution with a broad energy-dependent instru-
mental psf. The fact that the psf varies with position in
the image will be neglected in the following examples, be-
cause equation (1.1) is inapplicable in such situations and,
if the bulk of the emission is very concentrated then the
instrumental psf will be approximately constant over the
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
region of interest, in which case this treatment would be
valid. ASCA has a particularly broad psf for an X-ray in-
strument, so its use for imaging might appear rather sur-
prising. However, the good spectral resolution, coupled with
the energy-dependence of the psf creates a situation where
an image reconstruction algorithm could, for instance, be
usefully employed in determining temperature maps of the
ionised gas in X-ray clusters. Poisson, rather than Gaussian,
noise is appropriate for these images where the numbers of
photons per pixel is small.
10
V. Eke
3.2.1 Data production
Two different surface brightness profiles were created in a
2562 pixel grid according to
Σ(r) =
Σ0
(1 + (r/rc)2)3β/2 + b.
(3.3)
This is the β-profile proposed by Cavaliere & Fusco-Femiano
(1976) to represent cluster X-ray surface brightness profiles
with b corresponding to an additional background contri-
bution. The high signal-to-noise model had (Σ0, rc, β, b) =
(320 counts/pixel, 5 pixels, 0.7, 0.1 counts/pixel) whereas the
low signal-to-noise truth used (7, 6, 0.7, 0.001). These truths
were chosen to be similar to what the ASCA satellite would
have seen when looking at a cluster in the energy ranges
1− 2 keV and 7− 8 keV. A non circularly symmetric ASCA-
like psf having a full width half maximum of ∼ 10 pixels was
applied to these two truths and 10 Monte Carlo realisations
of the resulting T ∗ Bs were made. After smearing out with
the beam, the maximum counts per pixel were ∼ 70 and 1.5
for the two data sets, giving signal-to-noise ratios of √70
and √1.5 at the peak of the emission.
3.2.2 Results
Reconstructions were performed using 12 pixons ranging in
size from δ1 ≈ 1 pixel to 3δnpixon ≈ 100 (separated by fac-
tors of ∼ 1.3), ψ = 0.8 and υ = 0.6. These choices were kept
the same for both the data sets. For the Poisson distributed
noise the expected amplitude of the noise in pixel x was set
such that σ(x)2 = ( T ∗ B)(x). The comparison ML recon-
structions were performed by setting all pixon widths across
the pseudoimage to equal one pixel.
For the low signal-to-noise example, a tolerable fit ac-
γ , could actually be obtained with a flat T .
cording to χ2
Only when the misfit statistic was changed to ER was it
necessary to insert a source into the inferred truth in or-
der to produce a good fit. That is, the correlated residuals
produced when a uniform T was used to describe the weak
source were sufficiently small that their amplitudes were sta-
tistically acceptable. However the spatial correlation of the
residuals did have the power to discriminate between this
residual field and the anticipated noise.
Figure 4 shows the central regions of the 'X-ray cluster'
for the high signal-to-noise data. The top-left panel shows
the truth, and one of the realisations of the observed data
is shown beneath this. Both the smearing out of the sharply
peaked emission and the introduction of noise are very evi-
dent. The pixon-inferred truth for this particular D is shown
in the top-right panel and the corresponding ML-inferred
truth is contained in the final panel. It can be seen that the
noise is greatly suppressed by the pixon method and much
of the peaked emission has been recovered on sub-psf scales.
The ML reconstruction also removes noise in the central re-
gions, but at large radii spurious features are introduced,
essentially fitting to the noise. Also, at small radii, the ML
deconvolution does not do a good job of recovering the un-
smeared profile.
More detailed results concerning the radial surface
brightness profiles are shown in Figure 5, for both the high
and low signal-to-noise data sets. The mean pixon-inferred
profiles are drawn as dotted lines along with error bars show-
ing the standard deviation of the individual Monte Carlo re-
constructions. Long dashed lines represent the correspond-
ing ML quantities. It is apparent and reassuring, particularly
for the high signal-to-noise data, that the pixon algorithm
tends to produce similar profiles, independent of what noise
realisation has been used. The ML results show a signifi-
cantly larger dispersion in the reconstructions arising from
the different noise realisations. However, for the pixon re-
constructions, the error bars are sufficiently small that the
systematic deviations between the inferred and actual truths
can be seen. These are produced by the lack of richness in
the pixon description which leads to a preference for partic-
ular profiles. Nevertheless it is encouraging that the speedy
pixon algorithm can yield such results for two very differ-
ent quality data sets. In addition, the suppression of noise
is very effective, with no hint of any spurious sources being
produced in any of the realisations even for the low signal-
to-noise simulations, in contrast to the ML reconstructions.
4 CONCLUSIONS
The details of a speedy pixon method for image reconstruc-
tion have been given. This algorithm reduces the run time
from the n2 in the original procedure described by PP93
to nlogn, making the treatment of 2562 pixel images possi-
ble using only a few minutes on a typical workstation. The
application of the method to two types of simulated data
sets shows its ability to detect sources in low signal-to-noise
data without introducing spurious objects, as well as de-
convolving the instrumental psf. These results are a marked
improvement over a simple maximum likelihood reconstruc-
tion procedure which is applied in the data pixel grid and
includes a uniform image prior term. A more detailed study
of the ability of the pixon method to find clusters through
their S-Z distortion of CMB maps is in progress, including
a comparison with MEM.
ACKNOWLEDGMENTS
I would like to thank George Efstathiou for his helpful com-
ments, including pointing out to me the existence of pixons,
Rudiger Kneissl for his CMB map and many useful discus-
sions, David White for providing ASCA details and help-
ful suggestions, Shaun Cole for FFT assistance and Doug
Burke, Ofer Lahav and Radek Stompor for general enlight-
enment. This work was carried out with the support of a
PPARC postdoctoral fellowship.
REFERENCES
Bunn E., Fisher K.B., Hoffman Y., Lahav O., Silk J., Zaroubi S.,
1994, ApJ, 432, L75
Cavaliere A., Fusco-Femiano R., 1976, A&A, 49, 137
Dixon D.D. et al., 1996, A&AS, 120, 683
Dixon D.D. et al., 1997, ApJ, 484, 891
Eke V.R., Navarro J.F., Frenk C.S., 1998, ApJ, 503, 569
Fisher K.B., Lahav O., Hoffman Y., Lynden-Bell D., Zaroubi S.,
1995, MNRAS, 272, 885
Gull S.F., 1989,
in Skilling J., ed., Maximum Entropy and
Bayesian Methods. Kluwer, Dordrecht, p. 53
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Speedy pixon image reconstruction
11
ph/9901063)
Press W.H., Schechter P., 1974, ApJ, 187, 425
Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P.,
1992, Numerical Recipes, Cambridge University Press
Rybicki G.B., Press W.H., 1992, ApJ, 398, 169
Seljak U., Zaldarriaga M., 1996, ApJ, 469, 437
Skilling J., 1989,
in Skilling J., ed., Maximum Entropy and
Bayesian Methods. Kluwer, Dordrecht, p. 45
Slezak E., Bijaoui A., Mars G., 1990, A&A, 227, 301
Smith C.H., Aitken D.K., Moore T.J.T., Roche P.F., Puetter
R.C., Pina R.K., 1995, MNRAS, 273, 354
Sunyaev R.A., Zel'dovich Ya. B., 1972, Comm. Astrophys. Space
Phys., 4, 173
Tauber J., Pace O., Volont´e S., 1994, ESAJ, 18, 239
Weiner N., 1949, in Extrapolation and Smoothing of Stationary
Time Series. Wiley, New York
APPENDIX A: CONJUGATE GRADIENT
MINIMISATION DETAILS
The vast majority of the run time of the speedy pixon
method is spent calculating the derivative of the mis-
fit statistic with respect to the transformed pseudoim-
age values, Ht, and evaluating the inferred truth for
a given Ht. A couple of simple changes to the Nu-
linmin, sig-
merical Recipes line minimisation routine,
nificantly reduce the number of
function and deriva-
tive calls, and thus merit a mention here. (A more de-
tailed discussion of some of these issues is contained at
http://wol.ra.phy.cam.ac.uk/mackay/c/macopt.html.)
Firstly, the default tolerance requested by linmin is ∼ 103
times more stringent than need be for this application. Sec-
ondly, the initial guesses at bracketing the step size required
to reach the function minimum along the chosen direction
can be made more efficiently. Rather than keeping them
fixed at 0 and 1, these sizes should reflect the fact that the
different steps in parameter space are likely to have simi-
lar magnitudes. Thus the initially guessed step size should
be inversely proportional to the modulus of the vector along
which the step is to be taken. In addition, allowing these esti-
mates to adapt to previous values also leads to a more rapid
minimisation. Tuning the routine along these lines leads to
a speed up of about an order of magnitude for the examples
considered in this paper.
APPENDIX B: CALCULATION OF THE
MISFIT STATISTIC DERIVATIVES
The conjugate gradient misfit statistic minimisation requires
the evaluation of the partial derivatives of the statistic (here-
after labelled µ) with respect to the transformed pseudoim-
age variables. The chain rule for differentiation gives
∂µ
∂µ
=
∂H(x)
∂Ht(x)
.
(B1)
∂H(x)
∂Ht(x)
Defining the partial derivative of µ with respect to ( T ∗B)(x)
by F (x), the derivative of µ with respect to the pseudoimage
can be written as
∂µ
∂H(x)
=
npixon
Xl=1
W (δ(x), δl)[((F ⊗ B) × Vl) ⊗ Pl](x).
(B2)
Figure 5. Azimuthally averaged profiles, showing the perfor-
mance of the pixon (dotted lines) and ML (long-dashed lines)
algorithms for reconstructing high (upper panel) and low signal-
to-noise beta profiles. Error bars on the reconstructed results rep-
resent the standard deviation of 10 Monte Carlo realisations of
the same truth. For clarity, the error bars for the ML results have
been displaced 0.02 to the right. The full width half maximum of
the beam is ∼ 10 pixels. Solid and short-dashed lines represent
the true β profiles and the data respectively.
Hobson M.P., Jones A.W., Lasenby A.N., Bouchet F.R., 1998,
MNRAS, 300, 1
Knodlseder J. et al., 1996, SPIE Vol. 2806, 386
Lahav O., Fisher K.B., Hoffman Y., Scharf C., Zaroubi S., 1994,
ApJ, 423, L93
Metcalf T.R., Hudson H.S., Kosugi T., Puetter R.C., Pina R.K.,
1996, ApJ, 466, 585
Mighell K.J., 1999, ApJ, 518, 380
Pina R.K., Puetter R.C., 1992, PASP, 104, 1096 (PP92)
Pina R.K., Puetter R.C., 1993, PASP, 105, 630 (PP93)
Puetter R.C., 1996, SPIE Vol. 2827
Puetter R.C., Yahil A., 1999, to appear in Mehringer D., Plante
R., & Roberts D., eds. ASP Conf. Ser., Astronomical Data
Analysis Software and Systems VIII. San Francisco (astro-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
12
V. Eke
γ respectively.
Vl represents a mask that is unity in pixels with δ = δl
and zero otherwise. The calculation can be seen to be a
series of npixon correlations, hence the nlogn scaling. F is
readily shown to be −2R/(nσ2) and −2(D + min(D, 1)− T ∗
B)/(n(D + 1)) for χ2 and χ2
In the case of the autocorrelation of the residuals when
the anticipated noise amplitude is independent of signal and
position in the image (i.e. for the examples considered in Sec-
tion 3.1), equation (B2) is still applicable, but the derivative
of ER with respect to ( T ∗B)(x) needs to include a sum over
lag terms:
F (x) = −2
AR(z)(R(x + z) + R(x − z)).
nσ2 Xz
(B3)
The sum of residuals comes from the derivative of AR(z)
with respect to R(x).
When the anticipated noise amplitude σ also depends
upon the signal in the pixel, as is the case for the Poisson
noise used in the examples in Section 3.2, the derivative cal-
culation becomes somewhat more involved. Referring back
to equation (B2), the required form for F becomes
F (x) = −2
AR(z)K(x)(L(x + z) + L(x − z)).
(B4)
n Xz
where
K(x) =
and
L(x) =
D(x) − 0.5R(x)
( T ∗ B)(x)1.5
R(x)
p( T ∗ B)(x)
.
(B5)
(B6)
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0601576 | 3 | 0601 | 2006-04-18T16:21:02 | Cosmic Microwave Background Polarization | [
"astro-ph"
] | Cosmic microwave background (CMB) anisotropy is our richest source of cosmological information; the standard cosmological model was largely established thanks to study of the temperature anisotropies. By the end of the decade, the Planck satellite will close this important chapter and move us deeper into the new frontier of polarization measurements. Numerous ground--based and balloon--borne experiments are already forging into this new territory. Besides providing new and independent information on the primordial density perturbations and cosmological parameters, polarization measurements offer the potential to detect primordial gravity waves, constrain dark energy and measure the neutrino mass scale. A vigorous experimental program is underway worldwide and heading towards a new satellite mission dedicated to CMB polarization. | astro-ph | astro-ph |
Cosmic Microwave Background Polarization
James G. Bartlett
APC, 11 pl. Marcelin Berthelot, 75231 Paris Cedex 05, FRANCE
(UMR 7164 CNRS, Universit´e Paris 7, CEA, Observatoire de Paris)
E-mail: [email protected]
Abstract. Cosmic microwave background (CMB) anisotropy is our richest source of
cosmological
information; the standard cosmological model was largely established thanks
to study of the temperature anisotropies. By the end of the decade, the Planck satellite
will close this important chapter and move us deeper into the new frontier of polarization
measurements. Numerous ground -- based and balloon -- borne experiments are already forging into
this new territory. Besides providing new and independent information on the primordial density
perturbations and cosmological parameters, polarization measurements offer the potential to
detect primordial gravity waves, constrain dark energy and measure the neutrino mass scale.
A vigorous experimental program is underway worldwide and heading towards a new satellite
mission dedicated to CMB polarization.
1. Introduction
Observations of the cosmic microwave background (CMB) anisotropy have driven the remarkable
advance of cosmology over the past decade [1]. They tell us that we live in a spatially flat
universe where structures form by the gravitational evolution of nearly scale invariant, adiabatic
perturbations in a predominant form of non -- baryonic cold dark matter. Together with either
results from supernovae Ia (SNIa) distance measurements [2], the determination of the Hubble
constant [3] or measures of large scale structure [4], they furthermore demonstrate that a
mysterious dark energy (cosmological constant, vacuum energy, quintessence...) dominates the
total energy density of our Universe. These observations have established what is routinely called
the standard cosmological model: ΩM ≈ 0.3 = 1−ΩΛ, ΩBh2 ≈ 0.024 and H0 ≈ 70 km/s/Mpc [5].
Because the observations in fact over -- constrain the model, they test its coherence and its
foundations, marking a new era in cosmology.
The CMB results are remarkable for several reasons. They show us density perturbations
on superhorizon scales at decoupling and therefore evidence for new physics (inflation or other)
working in the early universe. The observed peaks in the power spectrum confirm the key
idea that coherent density perturbations enter the horizon and begin to oscillate as acoustic
waves in the primordial plasma prior to recombination; their position justifies the long -- standing
theoretical preference for flat space with zero curvature. Their heights measure both the total
matter and baryonic matter densities, and thereby provide direct evidence that most of the
matter is non -- baryonic; and, in a scientific tour de force, the CMB -- determined baryon density
broadly agrees with the totally independent estimation from Big Bang Nucleosynthesis [6].
These milestones were all obtained from study of the temperature, or total
intensity,
anisotropies. The Planck mission1 (launch 2007/2008) will largely complete this work by decade's
end with foreground -- limited temperature maps down to ∼ 5 arcmin resolution, leaving only the
smallest scales unexplored. With this in mind, the field is already turning to CMB polarization
measurements and their wealth of new information.
2. Polarization
Thomson scattering generates CMB polarization anisotropy at decoupling [7]. This arises from
the polarization dependence of the differential cross section: dσ/dΩ ∝ ǫ′ · ǫ2, where ǫ and
ǫ′ are the incoming and outgoing polarization states [8]; only linear polarization is involved.
This dependence means that an observer measuring a given polarization sees light scattered
preferentially from certain directions around the scattering electron in the last scatter surface.
The orthogonal polarization preferentially samples different parts of the sky. Any local intensity
anisotropy around the scattering electron thus creates a net linear polarization at the observer's
detector. Quantitatively, only a local quadrupolar temperature anisotropy produces a net
polarization, because of the cos2 θ dependence of the cross section. Also notice that the signal
is generated in the last scattering surface, where the optical depth transits from large to small
values; the optical depth must of course be non -- zero, but too large a value would erase any local
anisotropy.
2.1. Describing CMB Polarization
Polarized light is commonly described with the Stokes parameters [8]. As we have seen, the
CMB is linearly polarized, so we use the Stokes parameters Q and U , each of which is defined
as the intensity difference between two orthogonal polarization directions. Let (x, y) and (x′, y′)
refer to two coordinate systems situated perpendicular to the light propagation direction and
rotated by 45 degrees with respect to each other. Then Q ≡ Iy − Ix and U ≡ Iy′ − Ix′.
Clearly, the values of Q and U will depend on the orientation of the coordinate system used
at each point on the sky. Although from an experimental viewpoint this is unavoidable, it is
better for theoretical purposes to look for a coordinate -- free description; this latter could then
be translated into any chosen coordinate system. Two such descriptions were first proposed for
the CMB by Zaldariagga & Seljak and by Kamionkowski et al. [9]. The former, in particular,
model polarization as a spin 2 field on the sphere, an approach used in the publically available
CMB codes [10].
The coordinate -- free description distinguishes two kinds of polarization pattern on the sky by
their different parities. In the spinor approach, the even parity pattern is called the E -- mode
and the odd parity pattern the B -- mode. We can represent the polarized CMB sky by a map
color -- coded for intensity and with small bars indicating the direction of linear polarization at
each point. Consider a peak in the intensity (see Figure 1). If the polarization bars are oriented
either in a tangential or a radial pattern around the peak, we have a E -- mode; if they are oriented
at 45 degrees (relative to rays emanating from the peak), we have B -- mode: a reflection of the
sky about any line through the peak leaves the E -- mode unchanged (even parity), while the
B -- mode changes sign (odd parity)2.
Another useful way to see this is to consider the wave vectors of the plane wave perturbations
making up the intensity peak; they radially point towards the peak center (see Figure). We then
see that an E -- mode plane wave has its polarization either perpendicular or parallel to the wave
vector. A B -- mode plane wave, on the other hand, has a linear polarization at 45 degrees to the
wave vector. The wave vector in fact defines a natural coordinate system for definition of the
1 http://www.esa.int/science/planck
2 These local considerations generalize to the sphere [9].
k
k
Figure 1. Polarization patterns around an
local
intensity extremum. The upper row
shows the even parity E -- mode (negative on
the left, positive on the right), and the lower
row the odd parity B -- mode (negative on
the left, positive on the right). The red
arrows illustrate the projected wave vectors
of plane wave perturbations converging at the
extremum.
2.
Angular
power
the
temperature power
spectra
Figure
The bold solid black line
l(l + 1)Cl/2π.
spectrum
shows
of the standard model
[5]; the thin black
line gives the tensor contribution to the
temperature power for r = 0.5. The green
(upper) and blue (lower) short dashed curves
are, respectively, the scalar T E (absolute
value shown) and EE power spectra for
the standard model;
is well
measured on large scales by WMAP [24].
The red long dashed lines indicate the tensor
B -- mode power for r = 0.5 (upper) and
r = 10−4 (lower). Gravitational
lensing
produces the B -- mode power shown as the
red 3 -- dot -- dashed curve peaking at l ∼ 1000.
the former
Stokes parameters: in this system, Q=E and U =B. This is particularly useful when discussing
interferometric observations.
2.1.1. The E -- B Decomposition This decomposition of polarization into E and B -- modes is
powerful and practical. As we shall see below, the two different modes are generated by
different physical mechanisms, which is not surprising, since they are distinguished by their
parity. Secondly, their different parity also guaranties that we can separate and individually
measure the two modes and the total intensity on the sky. This is extremely important because
the intensity and polarization anisotropies have very different amplitudes (see Figure 2).
Theory predicts that the primary CMB anisotropy is a Gaussian field (of zero mean), and
current observations remain fully consistent with this expectation. We therefore describe CMB
anisotropy with the power spectrum Cl, which is nothing other than the second moment of the
field in harmonic space (i.e., the variance). As stated, most of the CMB milestones have been
obtained from temperature measurements, which in this context means from measurement of
the temperature angular power spectrum C TT
. With the introduction of polarization, we see
l
that in fact there are a total of 4 power spectra to determine: C TT, C TE, C EE, C BB; parity
considerations eliminate the two other possible power spectra3, C T B and C EB.
2.2. The Physical Content of CMB Polarization
In the standard model, inflation generates both scalar (S), or density perturbations [11] and
tensor (T), or gravity wave perturbations [12]. The scalar perturbations are created by quantum
fluctuations in the particle field (usually assumed to be a scalar field) driving inflation. After
inflation, these perturbations grow by gravity to form galaxies and the observed large scale
structure. Gravity waves, on the other hand, decay once they enter the horizon, and thus leave
their imprint in the CMB on large angular scales (around and larger than the decoupling horizon,
∼ 1 deg).
Gravity wave production by inflation, although a reasonable extrapolation of known physics,
would nevertheless be something fundamentally new. These waves would not be generated
by any classical or even quantum source (i.e., by the right -- hand -- side of Einstein's equations);
we suppose instead that the gravitational field itself (more specifically, the two independent
polarization states of a free gravity wave in flat space) experiences vacuum quantum fluctuations
like a scalar field. Finding gravity waves from inflation would therefore not simply be a detection
of gravity waves, but also a remarkable observation of the semi -- classical behavior of gravity.
Both scalar and tensor perturbations contribute to the temperature power spectrum;
in
practice, however, the scalar mode dominates, so that the measured temperature spectrum
effectively fixes the scalar perturbation amplitude. We quantify the relative amplitude of the
scalar and tensor perturbations by the parameter r ≡ PT/PS, where PT and PS represent the
power in the respective modes at a pivot wavenumber4 [13]. Since the scalar power is measured,
we use r to express the gravity wave amplitude. The WMAP data combined with large scale
structure data limit r < 0.53 (95% [14]). Unlike the scalar modes, which depend on the slope of
the inflation potential, the gravity wave amplitude depends only on the energy scale of inflation,
EI (specifically, PT ∝ (EI/Mpl)4, where Mpl is the Planck mass). Quantitatively5, we have
EI = 3.4 × 1016 GeV r1/4. Thus, the above limit on r corresponds to an upper limit on the
inflation scale of EI < 2.9 × 1016 GeV.
In addition to these primary CMB polarization signals, gravitational lensing by structures
forming along the line -- of -- sight to the last scattering surface sources a particular kind of
secondary polarization signal; I discuss this in the following section.
Figure 2 shows inflationary predictions for the various CMB power spectra from inflation --
generated scalar and tensor perturbations6. The temperature power spectrum for the standard
cosmological model (fit to the WMAP and other higher resolution experimental data) is given
by the bold solid black curve. This is now measured to the cosmic variance limit out to the
second peak. The light, black solid curve gives the maximum allowable tensor contribution to
the temperature power spectrum, i.e., at the current limit of r < 0.53. Note that it would be
very hard to significantly improve on this limit with only temperature measurements. This is
why there is such intense interest in polarization.
the E -- B Decomposition Although both scalar and tensor
2.2.1.
perturbations generate temperature anisotropy (C TT), primordial scalar perturbations only
The Importance of
3 The cosmological principle prohibits any preferred parity in the clustering hierarchy, implying that statistical
measures of the primordial anisotropy have even parity.
4 In [14], they use k = 0.002 Mpc−1.
5 The numerical relation refers to the definition of r used in the above references and corresponds to the parameter
employed in the CAMB code; it furthermore adopts the 2σ upper limit on the scalar power amplitude given by
WMAP.
6 These calculations where made using CAMB (http://www.camb.info)
produce E -- mode polarization, and hence only contribute to C EE and C TE. They cannot create
B -- mode polarization. We see this by considering a plane wave scalar perturbation passing
over a scattering electron: the local intensity quadrupole around the electron must be aligned
with the wave vector, which implies that the polarization of the scattered light must be either
perpendicular or parallel to the projected wave vector -- in other words, a pure E mode. The
axial symmetry imposed by the scalar nature of the density perturbations prevents any B -- mode
production.
The short dashed green (upper) and blue (lower) lines show C TE and C EE power spectra
from scalar perturbations for the standard model. These predictions follow directly from the
measured temperature power spectrum and the assumption -- usually adopted in the standard
model -- that the scalar perturbations are purely adiabatic. Given the measured temperature
spectrum, we could change the predicted C TE and C EE spectra by adding isocurvature
perturbations. Observations of these polarization modes will therefore constrain the presence of
such isocurvature modes7.
The TE cross spectrum in fact changes sign, but since I only plot its absolute value, the curve
oscillates between the sharp dips corresponding to the sign change. The additional bump at low
multipole arises from reinonization, for which I have taken an optical depth of τ = 0.17 [5].
In contrast to scalar perturbations, gravity waves (T) push and pull matter in directions
perpendicular to their propagation, aligning the local
intensity quadrupole in the plane
perpendicular to the wave vector. The loss of axial symmetry allows both E and B -- mode
production. Since the expansion dampens gravity waves on scales smaller than the horizon,
these tensor effects only appear on angular scales larger than ∼ 1◦ (the angular size of the
decoupling horizon). Hence, B -- mode polarization on large angular scales is the unique signature
of primordial gravity waves (the so -- called smoking gun) [15].
As mentioned, the amplitude of the gravity wave signal depends only on the energy scale of
inflation. A measurement of B -- mode polarization on large scales would give us this amplitude,
and hence a direct determination of the energy scale of inflation.8. The red long -- dashed curves
in Figure 2 show the tensor B -- mode spectrum for two different amplitudes -- the upper curve
for the the current limit of r < 0.5 (EI ∼ 3 × 1016 GeV), and the lower one for r = 10−4
(EI ∼ 3.4 × 1015 GeV).
Gravitational lensing of CMB anisotropy by structures forming along the line -- of -- sight to
decoupling also generates B -- mode polarization, but on smaller scales [16]. Lensing deviates
the photon trajectories (preserving surface brightness) and scrambles (distorts) our view of
the decoupling surface [17]. The E -- B modes are defined as pure parity patterns on the sky;
scrambling any such pattern will clearly destroy its pure parity, thereby leaking power into the
opposite parity mode.
If, for example, there were only E -- mode perturbations at decoupling
(e.g., gravity waves are negligible), we would still see some B -- mode in our sky maps on small
angular scales caused by gravitational lensing.
The lensing signal has both negative and positive aspects in the present context. On the
it masks the gravity wave B -- mode with a foreground signal with an identical
down side,
electromagnetic spectrum; thus, we cannot remove it using frequency information. We can,
however, extract and remove the lensing signal by exploiting the unique mode -- mode coupling
(between different multipoles, absent in the primary anisotropies) induced by the lensing [18].
Uncertainty in this process may ultimately limit our sensitivity to gravity waves [19].
On the positive side, the lensing signal carries important information about the matter power
spectrum and its evolution over a range of redshift inaccessible by any other observation. This
provides us with a powerful means of constraining dark energy and a singular method for
determining the neutrino mass scale [20]. Since the expansion governs the matter perturbation
7 These modes are also constrained by large scale structure observations.
8 More precisely, at the end of inflation.
growth rate, comparison of the amplitude of the power spectrum at high redshift to its amplitude
today probes the influence of dark energy. The shape of the power spectrum, on the other hand,
is affected by the presence of massive neutrinos, which tend to smooth out perturbations by free
streaming out of over -- and under -- densities. The effect suppresses the power spectrum on small
scales.
Recent studies indicate that by measuring the lensing polarization signal to the cosmic
variance limit, we would obtain a 1σ sensitivity to the sum of the 3 neutrino masses of
σΣ = 0.035 eV [20]. This is extremely important: current neutrino oscillation data call
for a ∆m2 = (2.4+0.5
−0.6) × 10−3 eV2 (2σ) [21]9, implying that the summed mass of the three
neutrino species exceeds the ultimate CMB sensitivity. CMB polarization therefore provides a
powerful and unique way to measure the neutrino mass scale, down to values unattainable in the
laboratory.
The red triple -- dot -- dashed curve in Figure 2 shows the B -- mode polarization from lensing.
Its amplitude is set by the amplitude of the primordial E -- mode signal and of the matter power
spectrum as it evolves. Since gravity waves generate both E -- modes (the tensor contribution
is not shown in the figure) and B -- modes of roughly equal power, we expect the scalar E --
mode to dominate. Thus, we have a good idea of the overall amplitude of the lensing B -- mode
spectrum, although the exact amplitude and shape will depend, as discussed, on the presence of
isocurvature modes, neutrinos and the nature of dark energy. For the curve shown in Figure 2,
I have adopted the standard model (no isocurvature perturbations) with a pure cosmological
constant and have ignored neutrinos.
3. Observational Effort
We often refer to polarization as the next step in CMB science. While appropriate, this
erroneously gives the impression that it remains for the future, when in fact, different experiments
have already measured CMB polarization. I give a summary in Figures 3 and 4.
The DASI experiment at the South Pole was the first to detect CMB polarization, both E
and T E modes [22]; their recently published 3 -- year results [23] are shown in Figures 3 and 4 as
the green squares. The original DASI detection was followed by WMAP's measurement of C TE
on large scales down to l ∼ 500 from the first year data [24]; these are not reproduced here.
More recently, the BOOMERanG collaboration reports measurements of C TT, C TE and C EE
and a non -- detection of B -- modes [25]. Combining their new BOOMERanG data with other
CMB and large scale structure data, MacTavish et al. [26] constrain r < 0.36 (95%). These
results are shown in Figures 3 and 4 as the red diamonds.
The CBI experiment has also published new measurements of C TT, C TE and C EE, as well
as a non -- detection of B -- modes [27]. These are shown in Figures 3 and 4 as the blue triangles.
Finally, the black asterisk in Figure 4 gives the E -- mode power measurement by CAPMAP [28].
All of these results are consistent with each other and with the prediction of the standard
cosmological model assuming pure adiabatic modes, shown in the figures as the black curve.
Although these new B -- mode limits are still far from placing any important constraints on
either gravity waves or lensing, the results are significant for what they imply about Galactic
foregrounds. We expect these foregrounds to generate E -- and B -- modes with equal strength --
there is no symmetry preferring one over the other. The lack of B -- mode power thus suggests
that foreground contamination in these data sets is well below the measured E and T E signals.
Scheduled for launch in 2007/2008, the Planck satellite will greatly advance our knowledge of
CMB polarization by providing foreground/cosmic variance -- limited measurements of C TE and
C EE out beyond l ∼ 1000. We also expect to detect the lensing signal, although with relatively
9 Here, ∆m2 is the difference between the singlet neutrino mass squared and the mean squared mass of the
neutrino doublet; see reference.
Figure 3. TE power spectra l(l + 1)Cl/2π.
The curve shows the power predicted by the
standard model (and measured on large scales
by WMAP [24], although not reproduced
here). Red diamonds give the BOOMERanG
results, green boxes the DASI 3 -- year results
and blue triangles the CBI results. The thin
black error bars show the BOOMERanG T B
power, a foreground tracer.
Figure 4. EE and BB power spectra. The
curve shows the standard model prediction for
C EE. Points are labeled as for the previous
figure; the black asterisk is the CAPMAP
E -- mode power measurement.
the
thin black error bars give each experiment's
B -- mode measurements (all consistent with
zero).
Here,
low precision, and could see gravity waves at a level of r ∼ 0.1. The Planck blue book quantifies
these expectations.
A leap in instrument sensitivity is required in order to go beyond Planck and get at the B --
modes from lensing and gravity waves. This important science is motivating a vast effort world
wide at developing a new generation of instruments based on large detector arrays. Numerous
ground -- based and ballon -- borne experiments are actually observing or being prepared. In the
longer term future, both NASA (Beyond Einstein) and ESA (Cosmic Vision) have listed a
dedicated CMB polarization mission as a priority in the time frame 2015-2020. Such a mission
could reach the cosmic variance limit on the lensing power spectrum to measure the neutrino
mass scale and perhaps detect primordial gravity waves from inflation near the GUT scale. The
exciting journey has begun.
References
[1] Smoot G F et al. 1992 ApJ 396 L1
Lange A E et al. 2001 Phy. Rev. D 63 042001
Stompor R et al. 2001 ApJ 561 L7
Bennett C L et al. 2003 ApJS 148 1
Scott D and Smoot G 2006 The Review of Particle Physics (Preprint astro -- ph/0601307)
[2] Knop R A et al. 2003 ApJ 598 102
Riess A G et al. 2004 ApJ 607 665
Astier P et al. 2005 Preprint astro -- ph/0510447
[3] Freedman W L et al. 2001 ApJ 553 47
[4] Tegmark M et al. 2004 ApJ 606 702
Cole S et al. 2005 MNRAS 362 505
[5] Spergel D N et al. 2003 ApJS 148 175
Seljak U et al. 2005 Phys. Rev D 71 103515
Sanchez A G et al. MNRAS in press (Preprint astro -- ph/0507583)
[6] Steigman G 2005, Preprint astro -- ph/0511534
[7] Bond J R and Efstathiou G 1984 ApJ 285 L45
Hu W and White M 1997 New Astron. 2 323
[8] Rybicki G B and Lightman A P 1979 Radiative processes in astrophysics (New York: Wiley -- Interscience)
[9] Zaldarriaga M. and Seljak U 1997 Phys. Rev. D 55 1830
Kamionkowski A, Kosowsky A and Stebbins A 1997 Phys. Rev. D 55 7368
[10] Seljak U and Zaldarriaga M 1996 ApJ 469 437 (http://www.cmbfast.org)
Lewis A, Challinor A and Lasenby A 2000 ApJ 538 473 (http://camb.info)
[11] Mukhanov V F and Chibisov G V 1981 JETP Lett. 33 532
Guth A H and Pi S -- Y 1982 Phys. Rev. Lett 49 1110
Hawking S W 1982 Phys. Lett. B 115 295
Starobinsky A A 1982 Phys. Lett. B 117 175
Bardeen J M, Steinhardt P J and Turner M S 1983 Phys. Rev. D 28 679
[12] Starobinsky A A 1979 JETP Lett. 30 682
Veryaskin A V, Rubakov V A and Sazhin M V 1983 Soviet Astron. 27 16
Abbott L F and Wise M 1984 Nucl. Phys. B 244 541
Kamionkowski M and Kosowsky A 1999 Ann. Rev. Nucl. Part. Sci. 49 77
[13] Leach S M, Liddle A R, Martin J and Schwarz D J 2002 Phys. Rev. D 66 023515
[14] Peiris H V et al. 2003 ApJS 148 213
[15] Seljak U and Zaldarriaga M 1997 Phys. Rev. Lett. 78 2054
Kamionkowski M, Kosowsky A and Stebbins A 1997 Phys. Rev. Lett. 78 2058
[16] Zaldarriaga M and Seljak U 1998 Phys. Rev. D 58 023003
[17] Blanchard A and Schneider J 1987 A&A 184 1
Seljak U 1996 ApJ 463 1
Bernardeau F 1997 A&A 324 15
[18] Hu W and Okamoto T 2002 ApJ 574 566
Okamoto T and Hu W 2003 Phys. Rev. D 67 083002
[19] Knox L and Song Y-S 2002 Phys. Rev. Lett. 89 011303
Kesden M, Cooray A and Kamionkowski M 2002 Phys. Rev. Lett. 89 011304
Seljak U and Hirata C M 2004 Phys. Rev. D 69 043005
[20] Kaplinghat M, Lloyd K and Song Y -- S 2003 Phys. Rev. Lett. 91 241301
Kaplinghat M 2003 New. Astron. Rev. 47 893
Lesgourgues J, Perotto L, Pastor S and Piat M 2005 Preprint astro -- ph/0511735
[21] Fogli G L, Lisi E, Marrone A and Palazzo A 2005 Preprint hep -- ph/0506083
[22] Kovac J M, Leitch E M, Pryke C, Carlstrom J E, Halverson N W and Holzapfel W L 2002 Nature 420 772
[23] Leitch E M, Kovac J M, Halverson N W, Carlstrom J E, Pryke C and Smith M W E 2005 ApJ 624 10
[24] Kogut et al. 2003 ApJS 148 161
[25] Piacentini F et al. 2005 Preprint astro -- ph/0507507
Montroy T E et al. 2005 Preprint astro -- ph/0507514
[26] MacTavish C J et al. 2005 Preprint astro -- ph/0507503
[27] Sievers J L et al. 2005 Preprint astro -- ph/0509203
[28] Barkats D et al. 2005 ApJ 619 L127
|
astro-ph/0509591 | 1 | 0509 | 2005-09-20T18:18:19 | The Mass Spectra of Cores in Turbulent Molecular Clouds and Implications for the Initial Mass Function | [
"astro-ph"
] | We investigate the core mass distribution (CMD) resulting from numerical models of turbulent fragmentation of molecular clouds. In particular we study its dependence on the sonic root-mean-square Mach number $\Ms$. We analyze simulations with $\Ms$ ranging from 1 to 15 to show that, as $\Ms$ increases, the number of cores increases as well while their average mass decreases. This stems from the fact that high-Mach number flows produce many and strong shocks on intermediate to small spatial scales, leading to a highly-fragmented density structure. We also show that the CMD from purely turbulent fragmentation does not follow a single power-law, but it may be described by a function that changes continuously its shape, probably more similar to a log-normal function. The CMD in supersonic turbulent flows does not have a universal slope, and as consequence, cast some doubt on attempts to directly relate the CMD to a universal Initial Mass Function. | astro-ph | astro-ph |
Draft date: April 10, 2018
The Mass Spectra of Cores in Turbulent Molecular Clouds and
Implications for the Initial Mass Function
Javier Ballesteros-Paredes1, Adriana Gazol1, Jongsoo Kim2, Ralf S. Klessen3,
Anne-Katharina Jappsen3, and Epimenio Tejero1
1Centro de Radioastronom´ıa y Astrof´ısica, UNAM, Apdo. Postal 72-3 (Xangari), Morelia,
Michoac´an 58089, M´exico
j.ballesteros, [email protected]
2 Korea Astronomy and Space Science Institute, 61-1, Hwaam-Dong, Yuseong-Gu, Daejeon
305-348, Korea
[email protected]
3Astrophysikalisches Institut Potsdam, An der Sternwarte 16, 14482 Potsdam, Germany
[email protected]
ABSTRACT
We investigate the core mass distribution (CMD) resulting from numerical
models of turbulent fragmentation of molecular clouds. In particular we study
its dependence on the sonic root-mean-square Mach number Ms. We analyze
simulations with Ms ranging from 1 to 15 to show that, as Ms increases, the
number of cores increases as well while their average mass decreases. This stems
from the fact that high-Mach number flows produce many and strong shocks
on intermediate to small spatial scales, leading to a highly-fragmented density
structure. We also show that the CMD from purely turbulent fragmentation
does not follow a single power-law, but it may be described by a function that
changes continuously its shape, probably more similar to a log-normal function.
The CMD in supersonic turbulent flows does not have a universal slope, and
as consequence, cast some doubt on attempts to directly relate the CMD to a
universal Initial Mass Function.
Subject headings: stars:formation - turbulence
– 2 –
1.
Introduction
An isothermal supersonic shock with a Mach number Ms creates density enhancements
2, where ρ1 and ρ0 are the densities of the post- and pre-shocked gas (e.g.,
of ρ1/ρ0 = Ms
Spitzer 1978). Since molecular clouds are turbulent and supersonic, it can be expected
that their internal density structure is, at first order (i.e., neglecting gravitational or thermal
fragmentation), a direct consequence of the fragmentation by the chaotic, supersonic velocity
field (see, e.g., the reviews by V´azquez-Semadeni et al. 2000; Mac Low & Klessen 2004;
Scalo and Elmegreen 2004 and references therein), a process which has been called turbulent
fragmentation. Thus, it is reasonable to expect that supersonic turbulence plays a crucial
role in determining the mass distribution of dense cores. In fact, the gravoturbulent scenario
of star formation suggests that the cores are formed by compressible turbulent motions inside
molecular clouds and that some of those cores may become gravitationally unstable and form
stars, while others will redisperse in the ambient medium (Sasao 1973; Hunter & Fleck 1982;
Elmegreen 1993; Ballesteros-Paredes, V´azquez-Semadeni & Scalo 1999; Klessen et. al. 2000;
Padoan et al. 2001; Padoan & Nordlund 2002, hereafter PN02).
On the other hand, the mass distribution of young stars follows a well-known distribution
called the Initial Mass Function (IMF). For stellar masses M ≥ 1 M⊙ it shows a power-law
behavior dN/dlogM ∝ M Γ, with slope Γ = −1.3 (Salpeter 1955; Scalo 1998; Kroupa 2002;
Chabrier 2003). Understanding the origin of the IMF is one of the fundamental goals of a
complete theory of star formation. Although important progresses have been achieved on the
observational determination of the IMF, there are still several proposed models to explain
it (see reviews by Meyer et al. 2000; Mac Low & Klessen 2004, and references therein), and
there is no agreement in the community for a standard one. One of the more recent models
suggests that the IMF properties are a direct consequence of the core mass distribution
(CMD). Observational works (e.g., Motte et al. 1998; Testi & Sargent 1998) have reported
a slope of the high-mass wing of the dense core mass distribution (or mass spectrum) that is
similar to the slope of Salpeter's IMF, suggesting that those cores are the direct progenitors
of single stars. Since stars are born from dense cores this idea is, in principle, tempting.
However, there is a large number of physical processes that may play an important role
during the core fragmentation and the protostellar collapse (see., e.g., Klessen & Burkert
2000; Goodwin, Whitworth & Ward-Thompson 2004; Bate & Bonnell 2004). These make
it unclear whether a single core will give birth to one or more stars, and what determines
the masses of individual stars within a single core. Some of these processes are: (a) The
mass distribution of cores changes with time as cores merge with each other (e.g., Klessen
2001; Schmeja & Klessen 2004). (b) Cores generally produce not a single star but clusters
of stars, and so the relation between the masses of cores and those of individual stars is
unclear (e.g., Larson 1985, Hartmann 2001, Goodwin et al. 2004). In addition, there may be
– 3 –
(c) competitive accretion influencing the mass-growth history of individual stars (see, e.g.,
Bate & Bonnell 2004), (d) stellar feedback through winds and outflows, or (e) changes in
the equation of state introducing preferred mass scales (e.g., Scalo et al. 1998; Li, Klessen &
Mac Low 2003; Jappsen et al. 2005; Larson 2005).
Besides the uncertainties mentioned above, there are other important caveats when
looking for a direct relationship between the CMD and the IMF. For instance, even though
some observational and theoretical works for dense, compact cores fit power-laws in the high-
mass wing of the CMD, the actual shape of those CMDs is not necessarily a single-slope
power-law, but a function whose slope varies in a more continuous way, frequently similar to
a log-normal distribution. From a theoretical point of view, PN02 have argued that the mass
distribution of dense cores generated by turbulent fragmentation follows closely the Salpeter
distribution of intermediate- to high-mass newborn stars, with a slope of ∼ −1.3 and that
the slope of the CMD depends only on the slope of the turbulent energy spectrum. These
results have been taken as a proof that turbulent fragmentation is essential to the origin of
the stellar IMF. Recent numerical studies by Tilley & Pudritz (2004) and Li et al. (2004)
have reported that the prediction of the PN02 model for the CMD agrees with the results of
their simulations. We consider that some cautionary remarks concerning these theoretical
results are necessary. First, the dynamical range of the CMD where a power-law with slope
−1.3 is appropriate is usually smaller than one order of magnitude. Second, the reported
CMDs are calculated at one single epoch, instead of being averaged over several timesteps.
Due to the fact that stochastic fluctuations might be non-negligible in a single frame (see
§2.4), determining CMDs at one single time is prone to significant statistical fluctuations.
This is particularly relevant when the core statistics is small, as in Li et al. (2004). Third,
the behavior of the CMDs presented in those works seems to be a continuous change in the
slope of the CMD, from zero at the maximum of the histogram, to large negative slopes in
the high-mass range. In fact, Gammie et al. (2003) mentioned that the high-mass wing of
the core mass spectrum in their simulations has a slope that appears to be consistent with
the Salpeter law, although other non-power-law forms for the mass spectrum may well be
consistent. Finally, in these works there is no systematic study of the dependence of the
CMDs with the Mach number. For example, Tilley & Pudritz (2004) present simulations
with Mach numbers of 2 and 5, and indeed a close inspection of their Fig. 17 shows that the
CMDs are different in both models, even though they can be fitted with the power-law with
the slope of −1.3 in a small dynamical range.
As has been discussed by Klessen (2001) and by Schmeja & Klessen (2004), at a given
Mach number the shape of the mass spectrum of dense cores changes with the wavenumber
at which turbulence is driven. In a similar way, the slope of the mass spectrum may very
well vary with the root-mean-square (RMS) Mach number Ms of the flow, which is related to
– 4 –
the kinetic energy density of the system. In fact, using numerical simulations, Kim & Ryu
(2005, see also Cho & Lazarian 2004) had found that the density power spectra of isothermal,
turbulent flows, vary with the Mach number, becoming shallower when the Mach number is
increased. This flattening is the consequence of the dominant density structures of filaments
and sheets. This occurs because, in a supersonic turbulent flow, density fluctuations are built
up through a sequence of jumps, produced by a succession of new shocks within previously
shocked regions (V´azquez-Semadeni 1994; Passot & V´azquez-Semadeni 1998). Thus, when
increasing the Mach number, the time frequency and strength of shocks inducing density
enhancements also increase. Since the density power spectra becomes shallower (Kim &
Ryu 2005), this means that the flow contains smaller and denser structures. Thus, it can be
expected that the CMD depends on the Mach number.
In the present work we test this idea by using two different numerical schemes that
model the interior of molecular clouds. In our simulations we vary the value of Ms from 1
(trans-sonic turbulence) to 15 (supersonic, highly compressible turbulence), and show that
the form of the CMD does indeed depend on the Mach number. The plan of the paper is the
following: in §2 we briefly introduce the simulations used in the present work, and discuss
the method for finding clumps, showing that turbulent systems reach statistical equilibrium
after one turbulent crossing time. We also discuss, the importance of averaging the CMD
over several timesteps in order to erase statistical fluctuations. In §3 we present our results.
§4 discusses the application of the PN02 model to isothermal hydrodynamic turbulent fields
and the relationship between the core mass distribution and the IMF. Finally, in §5, we
summarize our results.
2. Simulations
In order to study how the mass spectra of the cores depend on the Mach number,
we analyzed two different sets of simulations made with two distinct numerical schemes.
The first set uses a Total Variation Diminishing (TVD) method, and the second one a
Smoothed Particle Hydrodynamics (SPH) approach. All the simulations presented here are
three-dimensional and isothermal. We concentrate on small subregions within a much larger
cloud. Therefore, periodic boundary conditions are adopted in both numerical schemes.
Since we are interested in the fragmentation produced by the sole presence of turbulence, the
simulations do not include self-gravity and magnetic fields. Being isothermal, the simulations
are scale-free, and the renormalization is somehow arbitrary, but in principle they reproduce
the behavior of molecular clouds of sizes ranging from less than 0.1 pc up to 10–20 pc.
Table 1 presents the properties of each run. In the first four columns we list the name of the
– 5 –
run, its RMS Mach number (average over several timesteps), the numerical method used,
and the resolution. The remaining columns will be introduced in §2.4.
2.1. Total Variation Diminishing
The TVD scheme is a second-order accurate upwind scheme whose implementation for
isothermal flows is described in detail by Kim et al. (1999). Here we just mention that
the simulations performed used a turbulence random driver following the method presented
in Stone, Ostriker & Gammie (1998). Each driving velocity component is generated in
Fourier space, and its power spectrum has a peak at a large scale (specifically at wavenumber
kfor = 2 (2π/L0), where L0 is the one-dimensional size of the computational box). The kinetic
energy input rate is adjusted in order to maintain a roughly constant RMS sonic Mach
number. The three simulations used in the present work have a resolution of 2563 cells, and
their three-dimensional RMS Mach numbers are 1, 3 and 6. An additional high-resolution
simulation (5123 cells) at Mach 3 has been performed, in order to check that the log-normal
type shape of the CMD is not an artificial result of the low-resolution simulation.
2.2. Smooth Particle Hydrodynamics
The SPH method is a Lagrangian scheme in which the fluid is represented by an ensemble
of particles, and flow quantities are obtained by averaging over an appropriate subset of
the SPH particles (Benz 1990). The method is able to resolve large density contrasts as
particles are free to move, and so naturally the particle concentration increases in high-
density regions. Thus, this method has spatially varying spatial resolution as a function of
density (see Ossenkopf, Klessen & Heitsch 2001). Details on the method can be found in
Klessen (2001; see also Klessen, Heitsch, & Mac Low 2000; and Klessen & Burkert 2000). In
the four SPH simulations analyzed in this paper the three-dimensional RMS Mach numbers
are 3.4, 5.7, 9 and 15. Again, we drive turbulence at large scales, but in this case in a narrow
range of wavenumbers kfor ≤ 2 (2π/L0).
As in the case of the TVD method, we performed an additional high resolution SPH
simulation (9,938,375 particles) at Mach 3.8, in order to check that the shape of the CMD
is not an artificial result of the resolution. In this case we used the parallel code GADGET
(Springel, Yoshida & White 2001), in which the same turbulent driving scheme as in Klessen
(2001) has been implemented (see Jappsen et al. 2005 for details).
– 6 –
2.3. Clumpfinding particularities
We analyze all the simulations over many equally spaced timesteps covering several
turbulent crossing times, τturb (where τturb is defined as by the one-dimensional size of our
computational box Lbox divided by the RMS velocity of the field –i.e., the Mach number).
To compute the CMD at each timestep, we use an adapted version of the Williams et al.
(1994) clumpfinding algorithm to find cores in the three-dimensional density data-cubes
calculated with the TVD scheme. This modified version allows us to find clumps even
in large (5123) density cubes. For the particle data, we used its equivalent clumpfinding
algorithm, developed by Klessen & Burkert (2000, see their Appendix A). In both schemes,
we used logarithmic contours separated by half order of magnitude. The parameters of the
clumpfinding algorithms in each data set are somehow different:
for the TVD runs, the
lower density threshold is 2 times the mean density, and for the SPH runs the lowest density
contour is 10−2.5 (∼ 1/50 the mean density). In both cases, the separation of contours is
logarithmic, but while in the TVD runs the increase of mass is in steps of half order of
magnitude, in the SPH we adopted the same procedure as Klessen (2001), having only 10
logarithmic isocontours starting from the lowest contour and increasing the exponent by 0.5.
Although the details of the CMD obtained with different parameters of the clumpfinding
algorithms may change (e.g., exact mass of each core), the trend of the CMD variation
with Mach number does not depend on the numerical method and/or the details of the
clumpfinding scheme (see §3). As the main purpose of this work is to observe this trend, the
conclusions we reach are also independent of those choices.
2.4. Temporal convergence considerations
Since the simulations start from arbitrary random initial conditions with density and
velocity physically decorrelated, it is convenient, first, to investigate the time at which the
fluid reaches statistical equilibrium. We have studied the temporal evolution of the energy
spectrum (not shown here for reasons of space), and found that after one turbulent crossing
time, the energy spectrum of each run does not vary substantially. To verify that, in addition
to the energy spectra, at this time the simulations have reached statistical equilibrium, we
show in Fig. 1 the Mach number as a function of time (upper panel), as well as the number of
cores in the SPH (middle panel) and TVD (lower panel) simulations. From this figure, where
solid lines represent the SPH data, and dotted lines represent the TVD data, it becomes clear
that, in fact, after 1 τturb, the RMS Mach number (upper panel) in the flow approaches to
– 7 –
a constant value1. The quoted numbers give the RMS Mach number Ms, averaged over all
frames for which t ≥ 1 τturb. Similarly, in the middle and lower panels, where the number of
cores is plotted as a function of time, it is again seen that the transients are given only for
the first turbulent crossing time. After this time, each frame may give statistical fluctuations
of the order of 20% in the number of cores.
An important observation from the middle and lower panels in Fig. 1 is that, given
the fact that important fluctuations in the number of cores are present, making conclusions
based on a CMD computed from individual frames is risky. To prevent statistical bias when
reporting the shape of the CMD, it is therefore important to calculate it as an average over
at least a few turbulent crossing times. It is convenient to note that, if the average were
calculated over a fraction of turbulent time, the statistical fluctuation of the small cores
will be erased, but the statistical bias of the larger cores will not, since in a fraction of one
turbulent time the small cores may change considerably, but the larger ones not.
Getting back to our Table 1, in its last four columns we show, for each simulation, the
number of frames used in the present analysis (column 5), the initial and final times (in units
of τturb) used in the analysis (columns 6 and 7), and the time separation between frames, in
units of turbulent crossing times (column 8).
3. Results
Figure 2 shows the time-averaged energy spectrum of each simulation. For comparison,
the expected slopes for an incompressible flow (β = −5/3) and for a shock-dominated one
(β = −2) are indicated by dashed straight lines. The kinetic energy density is proportional
to the total area below the curve. This area grows as the Ms increases. Note that the
energy spectra for the SPH runs have somewhat shallower slopes than those for the TVD
runs. However, this can be understood as a consequence of the considerable differences
between the two numerical schemes. The former is a particle-based method while the latter
uses finite differences on a equally-spaced mesh. Both methods are complementary (see also
the discussion in Klessen et al. 2000), and we base our discussion solely on trends that are
common in both sets of simulations. The fact that both numerical schemes, TVD and SPH,
give rise to the same behavior of the CMD with RMS Mach number (see §3), reassures us
1Note however, that Ms still shows variations of order 5% around the mean value, specially in the SPH
data. This is due to compressibility of the medium and the adopted fixed large-scale driving pattern, as
discussed by Klessen & Lin (2003). The deviations from the mean value between individual time frames
decrease with decreasing RMS Mach number.
– 8 –
to consider the observed effect of varying Ms as a real physical trend.
Figures 3a, b show snapshots of the density field in two TVD simulations with Ms =1 and
6, respectively. The snapshots have been acquired at t = 5 τturb (Ms = 1) and t = 4.5 τturb
(Ms = 6).
It is evident that the medium is more strongly fragmented in the high Mach
number case. A similar conclusion holds for the SPH data cubes. The interactions of
supersonic shocks and shocklets in high-RMS Mach flows produce strongly localized density
enhancements, and consequently give rise to high degrees of fragmentation even on scales
much smaller than the turbulent driving scale.
To quantify that further, we present in Figs. 4 and 5 the resulting core mass spectra
resulting from the TVD and the SPH models, respectively. The total mass in the computa-
tional box for all the models has been renormalized to 64 Jeans masses2. The CMD clearly
varies with Ms. In particular, the total number of cores as well as the peak value of the
histogram increase with Ms. To illustrate this result more clearly we show in Fig. 6 (a) the
total number of cores and, (b) the maximum value reached by the histograms as function
of Ms. In this figure, filled and open squares correspond to the SPH and the TVD models,
respectively. Solid lines in each panel are least-square fits. As a consequence of mass con-
servation (recall that each TVD and SPH simulation has the same total mass) the typical
core mass shrinks as the number of cores increases. This result is shown in Fig. 7, where we
plot (a) the mass of the most massive core and, (b) the mass at which the histogram peaks,
both as function of Ms.
A point of concern is whether our results depend on the resolution of the simulations.
In Fig. 8 we show the CMD resulting from the two high-resolution (HR) simulations. Recall
that the HR-TVD runs has 5123 pixels, while the HR-SPH runs has 9,938,375 particles. This
means that the (spatial) resolution is 8 times larger in the TVD-HR case, while the mass
resolution in the SPH-HR case is almost 50 times larger. From this figure we stress that the
shape of the CMD is, indeed, not a power-law function, but something more similar to a
log-normal function, for which the slope varies from zero at the maximum of the distribution
to large negative values.
Recall at this point that, since the clumpfinding algorithm in the SPH runs has been
2Even though we do not take the self-gravity of the gas into account, we give the total mass in units of
the Jeans mass. Other choices of the mass unit are possible and fully equivalent. The reason for quoting the
mass in Jeans masses is to allow for direct comparison with recent numerical works, e.g., by Klessen (2001),
Gammie et al. 2003, Tilley & Pudritz (2004); Li et al. (2004), as well as with a future contribution of the
analysis in the self-gravitating case. Note that we adopt a cubic definition of the Jeans mass, MJ = ρλ3.
This differs from the spherical definition, MJ = 4π/3ρ(λ/2)3, by a factor ∼ 2.
– 9 –
used with a smaller density threshold, the typical core mass in the SPH models exceeds that
in the TVD case. However, the general trend of increasing number of cores and decreasing
core mass with increasing RMS Mach number is independent of the used numerical method
as well as the details of the clumpfinding scheme.
4. Discussion
In this section we examine the implications of the results presented in §3, in the context
of previous work on the CMD and its relationship with the stellar IMF. We find that the
CMD is not a power-law and has a shape that depends on the RMS Mach number. This fact
contradicts previous results stating that, in the magnetic case, the CMD follows a power-law
whose slope depends on the power index of the kinetic energy spectrum (PN02). As has been
mentioned, our simulations do not include magnetic fields and this is an important difference
with the work of PN02. However, this fact is not the reason for the differences between the
shape of CMD reported in previous section and the power-law they obtain. Indeed, it can
be shown that a similar analysis to the one performed by PN02 but for the non magnetic
case, also leads to a power-law in which the RMS Mach number of the flow does not enter in
the functional form of N(m) d log m. It only determines the normalization factor, which, in
turn, implies that larger Ms leads to larger number of cores. Although this fact is consistent
with our numerical results (see §3), our results show that the actual shape of the CMD does
vary with Ms.
We identify the main reason why the model by PN02 does not include the dependency
of the shape of the CMD on Ms as hinted before: density fluctuations (cores) in a purely
turbulent fluid are built-up by a statistical superposition of shocks. This means that mass
and density contrast in individual cores grow by a sequence of density jumps with pre-shock
densities that can significantly deviate from the mean value ρ0 (see, e.g., Smith et al. 2000).
In addition, in a turbulent flow with characteristic Mach number Ms, the Mach numbers of
the individual shocks may deviate considerably from the mean value Ms. This translates
into a large scatter of the Larson (1981) δv − R relationship, which has been observed both
in theoretical works (e.g., V´azquez-Semadeni et al. 1997; Ostriker et al. 2001; Ballesteros-
Paredes & Mac Low 2002), as well as in works that include more recent sets of observations
includding different tracers (see Garay & Lizano 1999). Instead, the PN02 model and its
hydrodynamic pendant introduced above, assumes that all cores form in one single event
with fixed Mach number and initial density. Other caveat in the PN02 model is that stream
collisions form sheets, not cores. Thus, the size λ of the shock-bounded region may not be
a good representative of the size of cores.
– 10 –
If cores are generated not by single shocks, but by statistical superposition of subsequent
shocks of the turbulent flow, the dependency of the density structure on the RMS Mach
number may have important implications for the fragmentation behavior of molecular clouds,
clumps and their cores. For instance, the internal density structure of a dense core (even if it
is subsonic) may differ depending whether it is embedded in a highly supersonic, turbulent
molecular cloud, or in a more quiescent one (see, e.g., Ballesteros-Paredes et al. 2003, Klessen
et al. 2005). For high Mach numbers, the high- frequency part of the turbulent energy cascade
lies in the supersonic regime and, hence, is able to produce strong density fluctuations on
intermediate to small scales. The amount of energy at the largest scales, where most of
the power is, thus, should define the fragmentation on all scales, included those where, the
non-thermal motions became subsonic. Thus, we speculate that cores of a given mass in
more turbulent clouds (e.g., Orion) will have more substructure than cores of the same mass
in less turbulent clouds (e.g., Taurus).
As mentioned in the introduction, there have been various attempts to make a direct
connection between the CMD and the stellar IMF. However, the result found in the present
work that the CMD has an strong dependence on the Mach number implies that the rela-
tionship between these two distributions is not immediate. It is thus important to discus
about the possible reasons of this fact in the context of turbulent fragmentation.
The supersonic turbulent motions ubiquitously observed in interstellar clouds, suggest
that cores form in first place via shocks, with gravity taking over in the densest and most
massive regions. Once gas cores become gravitationally unstable, collapse sets in, but the
details of that collapse are not well understood, yet. It is tempting to suggest that turbulence
plays a critical role in determining the stellar IMF (see, e.g., the discussion in Ballesteros-
Paredes 2005). In this case, the IMF will directly follow from the CMD if individual molecular
cloud cores map one-to-one onto stars. However, there is a large number of physical processes
that may play an important role during protostellar collapse. Thus, the main problem
relating the CMD with the IMF is that a core with a given mass will not necessarily form
a star of the same mass, or of a mass that is proportional to the core mass. However, even
considering that these effects do not play an important role, and that every core of mass M
does indeed produce a star whose mass is directly proportional to M, the numerical results
presented in this paper suggest that turbulent fragmentation is not the sole agent defining
the shape of the IMF, since the shape of the CMD seems not to be a power-law and it varies
(as well as the CMD normalization) with the total amount of kinetic energy. Thus, although
turbulent fragmentation may play an important role in the determination of the CMD, it
is difficult to argue that it is the main ingredient for determining the IMF. In a supersonic
turbulent fluid, where the gas is compressed into filaments and sheets, the initial CMD may
depend on the parameters of the turbulent gas, however there is still room for other processes
– 11 –
to take over and modify the initial CMD in its way to an intermediate mass distribution,
and finally to the IMF.
Finally, it has to be mentioned that, although it is generally accepted that the IMF has
a slope of −1.3 at the high-mass range, the universal character of the IMF is still a debated
issue. For instance, Scalo (1998) notes that observations taken at face value reveal variations
in the IMF between different star clusters. Kroupa (2001; 2002), on the other hand, argues
that these variations can be explained by stellar dynamical processes during the early phases
of star-cluster evolution, like mass segregation and the preferred evaporation of low-mass
stars. On similar grounds, also Elmegreen (2000) argues in favor of a more universal IMF,
although he recognizes statistical deviations.
5. Conclusions
Using two different numerical schemes (TVD and SPH), we have shown that the core
mass distribution (CMD) or core mass spectrum from purely turbulent (i.e. without self-
gravity and magnetic fields) fragmentation follows a distribution with a continuously varying
slope, not necessarily a single power-law. The parameters of this distribution depend on the
kinetic energy density in the system as characterized by the RMS Mach number of the flow.
Interactions of shocks in high Ms flows are able to produce strong density fluctuations on
small and intermediate scales. We thus speculate that quiescent cores embedded in highly
supersonic molecular clouds should exhibit more internal structure than quiescent cores
embedded in more quiescent clouds. As consequence, for a molecular cloud of a given mas,
the resulting resulting CMD is characterized by a large number of low-mass cores when the
Mach number is high. Low-Ms flows, on the other hand, are characterized by small density
contrasts and the CMD features fewer cores with on average larger masses.
We conclude that the RMS Mach number of the turbulent flow is a very important
parameter that determines the shapes of CMDs. Our calculations indicate that the exact
shape of the CMD is not universal, but changes with the Mach number. This calls the
existence of a simple one-to-one mapping between the purely turbulent CMD and the ob-
served stellar initial mass function (IMF) into question. Finding the true relation between
CMD and IMF instead is a very complex and difficult task, and requires to take additional
physical processes into account, such as protostellar feedback from new-born stars (bipolar
outflows, winds, and radiation), subfragmentation of individual collapsing protostellar cores
into binary and low-N multiple systems, competitive accretion to name just a few.
We thank E. V´azquez-Semadeni and S. Kitsionas for fruitful discussions. This work
– 12 –
was supported in part by CONACYT grant 27752-E. The work by AG was supported by
the DGAPA grant PAPIIT-IN114802. J.S.K. was supported by the Astrophysical Research
Center for the Structure and Evolution of the Cosmos (ARCSEC) of the Korea Science and
Engineering Foundation through the Science Research Center (SRC) program. R.S.K. and
A.K.J. were supported by the Emmy Noether Program of the DFG under grant KL1358/1.
The TVD simulations were performed on the Linux clusters at KASI and CRyA, with funding
from KASI and ARCSEC, and CONACYT grant 27752-E, respectively. The SPH calcula-
tions have been performed at the University of California at Santa Cruz as well as the PC
cluster of the Astrophysical Institute Potsdam. This work has made extensive use of NASA's
Astrophysics Abstract Data Service and LANL's astro-ph archives.
Ballesteros-Paredes, J. 2005, Ap&SS, in press
REFERENCES
Ballesteros-Paredes, J., Klessen, R. S., & V´azquez-Semadeni, E. 2003, ApJ, 592, 188
Ballesteros-Paredes, J. & Mac Low, M. 2002, ApJ, 570, 734
Ballesteros-Paredes, J., V´azquez-Semadeni, E., & Scalo, J. 1999, ApJ, 515, 286
Bate, M. R., & Bonnell, I. A. 2004, MNRAS, 728
Benz, W. 1990, in The Numerical Modeling of Nonlinear Stellar Pulsations, ed. J. R. Buchler
(Dordrecht: Kluwer), 269
Chabrier, G. 2003, PASP, 115, 763
Elmegreen, B. G. 1993, ApJ, 419, L29
Elmegreen, B. G. 2000, ApJ, 539, 342
Gammie, C. F., Lin, Y., Stone, J. M., & Ostriker, E. C. 2003, ApJ, 592, 203
Garay, G., & Lizano, S. 1999, PASP, 111, 1049
Goodwin, S. P., Whitworth, A. P., & Ward-Thompson, D. 2004, A&A, 423, 169
Hartmann, L. 2001, AJ, 121, 1030
Hunter, J. H., & Fleck, R. C. 1982, ApJ, 256, 505
– 13 –
Jappsen A. K., Klessen R. S., Larson, R. B., Li Y., & Mac Low, M. M., 2005, A&A, 435,
611
Kim, J., Ryu, D., Jones, T. W., & Hong, S. S. 1999, ApJ, 514, 506
Kim, J.S. et al. 2005. ApJ, submitted
Klessen, R. S. 2001, ApJ, 556, 837
Klessen, R. S., Ballesteros-Paredes, J., V´azquez-Semadeni, E., & Dur´an-Rojas, C. 2004,
ApJ, 620,786
Klessen, R. S., & Burkert, A. 2000, ApJS, 128, 287
Klessen, R. S., Heitsch, F., & Mac Low, M. 2000, ApJ, 535, 887
Klessen, R. S., & Lin, D.N.C. PRE, 67, 036311
Kroupa, P. 2001, MNRAS, 322, 231
Kroupa, P. 2002, Astronomical Society of the Pacific Conference Series, 285, 86
Larson, R. B. 1981, MNRAS, 194, 809
Larson, R. B. 1985, MNRAS, 214, 379
Larson, R. B. 2005, MNRAS, 359, 211
Li, Y., Klessen, R. S., & Mac Low, M. M. 2003, ApJ, 592, 233L
Li, P. S., Norman, M. L., Mac Low, M., & Heitsch, F. 2004, ApJ, 605, 800
Mac Low, M., Klessen, R. S. 2004, RvMP, 76, 125
Meyer, M. R., Adams, F.C., Hillenbrand, L.A., Carpenter, J.M., & Larson, R.B. 2000, in
Protostars and Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russell (Tuscon:
Univ. Arizona Press), 121
Motte, F., Andr´e, P., & Neri, R. 1998, A&A, 336, 150
Ossenkopf, V., Klessen, R. S., & Heitsch, F. 2001, A&A, 379, 1005
Ostriker, E. C., Stone, J. M., & Gammie, C. F. 2001, ApJ, 546, 980
Padoan, P., Juvela, M., Goodman, A. A., & Nordlund, A. 2001b, ApJ, 553, 227
– 14 –
Padoan, P., & Nordlund, A. 2002, ApJ, 576, 870
Passot, T., & V´azquez-Semadeni, E. 1998, PhRvE, 58, 4501
Salpeter, E. E. 1955, ApJ, 121, 161
Sasao, T. 1973, PASJ, 25, 1
Scalo, J. 1998, in The Stellar Initial Mass Function (38th Herstmonceux Conference), G.
Gilmore & D. Howell eds. ASPC, 142, 201
Scalo, J., & Elmegreen, B. G. 2004, ARA&A, 24,275
Scalo, J., V´azquez-Semadeni, E., Chappell, D., & Passot T. 1998, ApJ, 504, 835
Schmeja, S., & Klessen, R. S. 2004, A&A, 419, 405
Smith, M. D., Mac Low, M.-M., & Heitsch, F. 2000, A&A, 362, 333
Spitzer, L. 1978. Physical processes in the interstellar medium, New York Wiley-Interscience
Springel, V., Yoshida, N., & White, S. D. M. 2001, New Astronomy, 6, 79
Stone, J. M., Ostriker, E. C., & Gammie, C. F. 1998, ApJ, 508, L99
Tilley, D. A., & Pudritz, R. E. 2004, MNRAS, 353, 769
Testi, L., & Sargent, A. I. 1998, ApJ, 508, 91
V´azquez-Semadeni, E. 1994, ApJ, 423, 681
V´azquez-Semadeni, E., Ballesteros-Paredes, J., & Rodriguez, L. F. 1997, ApJ, 474, 292
V´azquez-Semadeni, E., Ostriker, E. C., Passot, T., Gammie, C. F., & Stone, J. M. 2000,
Protostars and Planets IV, 3
Williams, J. P., de Geus, E. J., & Blitz, L. 1994, ApJ, 428, 693
This preprint was prepared with the AAS LATEX macros v5.2.
– 15 –
Fig. 1.- Temporal evolution of the Mach number (upper panel), number of cores found in
the SPH models (middle panel) and in the TVD models (lower panel). Solid lines denote
SPH runs, and dotted lines denote TVD runs. Quoted numbers in the upper panel give
the mean Mach number, averaged over the frames whose times are larger than 1 turbulent
crossing time. (see Table 1). Note that after 1 turbulent crossing time, the fluid has reached
equilibrium, since (a) the Mach number becomes nearly constant, with fluctuations not larger
than 6.5% (upper panel), and the number of cores has reached its typical value. Note that
still fluctuations up to 20% in the number of cores may be found, showing the importance
of averaging over several timesteps when calculating CMDs.
– 16 –
Fig. 2.- Energy spectrum for each simulation. The solid lines correspond to SPH simu-
lations and the dotted lines to TVD simulations. The dashed lines have a slope of −2 and
−5/3.
– 17 –
Table 1. Model properties.
Name
RMS Mach Number Method Resolutiona # frames
b
ti
b
tf
∆tframe
c
TVD1
TVD3
TVD6
SPH3
SPH6
SPH9
SPH15
TVD-HR
SPH-HRe
0.97
2.87
6.2
3.4
5.7
9.1
14.98
3
3.8
TVD
TVD
TVD
SPH
SPH
SPH
SPH
TVD
SPH
2563
2563
2563
205,000
205,000
205,000
205,000
5123
9,938,375
8
12
18
96
64
77
50
12
6
1.2
1.2
1.2
5
4.5
6
0.4
0.3
0.3
1.02
1.008
3.375
1.05
4.25
4.536
6.795
4.725
0.034
0.056
0.0405
0.075
1.2
4.5
1.368
1.558
0.3
0.038
aNumbers of cells for the TVD scheme, and particles for the SPH scheme.
bIn units of the turbulent crossing time.
dTime spacing between frames, in turbulent crossing timesteps.
eRun performed with Gadget.
Fig. 3.- Snapshots of the TVD runs with (a) Ms = 1 at t = 5 τturb, (b) Ms = 6 at
t = 4.5 τturb. Note how the turbulent fragmentation of the medium is different for different
Mach numbers. Blue, green, yellow, orange, and red colors represent isodensity surfaces with
1, 2, 3, 4 and 5 times the mean density in frame (a). In frame (b), the same order of colors
represents isodensity surfaces with 5, 11.25, 17.5, 23.75 and 30 times the mean value.
– 18 –
Fig. 4.- Time averaged core mass distribution resulting from TVD simulations. The two
dot-dashed lines with slopes of −0.5 and −1.3 are shown for the comparison purpose with
the mass functions of Giant Molecular Clouds and protostellar cores, respectively.
– 19 –
Fig. 5.- Time averaged core mass distribution resulting from SPH simulations. The two
dot-dashed lines with slopes of −0.5 and −1.3 are shown for the comparison purpose with
the mass functions of Giant Molecular Clouds and protostellar cores, respectively.
– 20 –
Fig. 6.- (a) Total number of cores, and (b) Number of cores at maximum of the distribution,
as a function of the mach number Ms. Filled (open) squares denote results from SPH (TVD)
runs. Solid lines represent least-square fits to the data. m represents the slope of the fit.
Note that, for each dataset, the number increases as Ms increases.
Fig. 7.- (a) Mass of the most massive core, and (b) mass at which the maximum of the
histogram occurs as a function of Ms. Filled (open) squares denote results from SPH (TVD)
runs. Solid lines represent least-square fits to the data. m represents the slope of the fit.
Note that, for each numeric scheme, the mass decreases as Ms number increases.
– 21 –
Fig. 8.- CMD for the high resolution runs using (a) the TVD, and (b) SPH schemes. Note
that even at high resolution, a single power-law not necessarily reproduces the high-mass
wing of the CMD. Instead, a function that changes more continuously its slope (probably a
log-normal function), from zero at maximum, to large negative values for increasing masses,
may reproduce better the distribution.
|
astro-ph/0112399 | 1 | 0112 | 2001-12-17T17:55:24 | The moon and the origin of life | [
"astro-ph"
] | Earth is unusual in bearing life, and in having a large moon. A number of authors have suggested a possible connection between the two, e.g. through lunar stabilisation of the earth's obliquity, or through the effects of the oceanic tides. The various suggestions are reviewed. | astro-ph | astro-ph |
The moon and the origin of life†
C.R. Benn
Isaac Newton Group, Apartado 321, 38700 Santa Cruz de La Palma, Spain
Abstract. Earth is unusual in bearing life, and in having a large moon. A number
of authors have suggested a possible connection between the two, e.g. through lunar
stabilisation of the earth's obliquity, or through the effects of the oceanic tides. The
various suggestions are reviewed.
Keywords: moon, origin of life
1. Introduction
The properties of the universe that we observe must be consistent with
the evolution of carbon-based life within it. This observational selec-
tion effect is known as the weak anthropic principle (Barrow & Tipler
1986). It has been invoked to explain a number of otherwise unlikely
coincidences such as the nuclear resonance that allows carbon to form
in stellar interiors, the big numbers coincidence (ratio of strengths
of electromagnetic and gravitational forces ∼ 1040 ∼ current size of
the observable universe in proton diameters), and, more recently, the
smallness of the cosmological constant (Efstathiou 1995) and the am-
plitude of the primordial density fluctuations which seeded the growth
of galaxies and clusters of galaxies (Tegmark & Rees 1998).
A similar selection effect will apply in our local astrophysical envi-
ronment. Clearly, the luminosity and lifetime of the sun, and the shape
and size of the earth's orbit, must be such as to maintain the earth's
surface, for a long period, at a temperature suitable for the evolution
of organic life (e.g. Kasting et al 1993). However, our solar system may
also be atypical in other respects:
• The sun's metallicity is unusually high for its age (Whittet 1997,
Gonzalez 1999), perhaps reflecting the higher probability of planetary
systems being associated with high-metallicity parent stars.
• The sun's luminosity may be unusually stable (Gonzalez 1999).
• Jupiter's role in ejecting comets from the solar system could have
been crucial in protecting the young earth from life-inhibiting impacts
(Wetherill 1995). Solar systems with Jupiter-like planets at similar radii
may thus not be typical.
• The sun may be orbiting the galaxy close to the co-rotation circle
† Earth, Moon and Planets 85/86, 61 (2001)
c(cid:13) 2018 Kluwer Academic Publishers. Printed in the Netherlands.
moon3.tex; 2/06/2018; 12:17; p.1
2
C.R. Benn
(Mishurov & Zenina 1999), which minimises the number of spiral-arm
crossings, and consequent disruption of the solar system (e.g. by nearby
supernovae, tidal effects).
Although it's possible that none of the above features of our solar
system is essential to the evolution of life on earth, the probability of
our observing them is enhanced if they increase the probability that
intelligent life will develop, i.e. it would not be surprising to observe
any feature F if its a priori probability, pF , satisfies:
pF > plif e(0)/plif e(F )
(1)
where plif e(0) and plif e(F ) are the probabilities of intelligent life
evolving respectively in the absence and in the presence of feature F .
Several authors (e.g. Butler 1980, Comins 1993) have suggested that
there might be a link between our large moon, arguably an a priori
unlikely feature of the earth's environment, and the evolution of life on
earth, i.e. that:
pmoon > plif e(0)/plif e(moon)
(2)
All 3 parameters in this inequality are unknown. Below we consider
what is known about the origin of the moon, and the origin of life, and
how the former might affect the latter.
2. The origin of the moon
The earth's moon is unusual amongst solar-system satellites in having
relatively high mass and in dominating the total angular momentum
of the earth-moon system. It is also unusually 'dry', lacking volatile
elements such as K and Bi. These characteristics, particularly the high
angular momentum, posed serious problems for the three main hy-
potheses for the origin of the moon until the mid 1980s: capture,
co-formation and fission (Hartmann et al 1985). There is now a con-
sensus (Taylor 1992) that none of these three hypotheses is tenable,
and that the earth-moon system probably formed when a planetesimal
∼ 0.1 - 0.2 earth masses underwent a grazing collision with the proto-
earth, soon after the formation of the solar system 4600 Myr ago. In this
widely-accepted 'Big Splash' scenario, the metallic core of the impact-
ing body sank to join the earth's, and some of the shattered mantle
reassembled in orbit to form the moon. Estimates of the probability
(pmoon) of the earth acquiring such a large moon have been boosted
by this new perspective, but e.g. Ringwood (1990) and Lissauer (1997)
argue that only a narrow range of initial conditions could have resulted
in a large moon in earth orbit.
moon3.tex; 2/06/2018; 12:17; p.2
The moon and the origin of life
3
3. The origin of life
Life on earth probably dates back at least 3500 Myr, i.e. to within a
few 100 Myr of the formation of the earth's crust 4200 - 4000 Myr
ago. The oceans at that time would have been a weak solution of
organic precursors (amino acids, pyrimidines etc.), formed in situ, or
in the comets that may have provided much of the water (Chyba 1990,
Lazcano-Araujo & Or´o 1981). How a self-replicating system evolved
from this primordial soup has been the subject of much speculation
and laboratory work, reviewed by Or´o et al (1990) and Deamer &
Fleischaker (1994). One common theme to emerge is the importance of
concentrating the soup to encourage polymerisation, e.g. in tidal pools
which repeatedly dry out under the sun, echoing Darwin's (1871) spec-
ulation about the origin of life in "some warm little pond with all sorts
of ammonia and phosphoric salts". For example, Or´o et al (1990) note
that the best contemporary laboratory models of pre-cellular systems
are liposomes (phospholipid vesicles), and encapsulation of DNA within
liposomes has been achieved by dehydration-hydration cycles similar to
those occurring in intertidal pools (Deamer & Barchfeld 1982).
Once self-replication is established, evolution to more complex sys-
tems can proceed through Darwinian selection. Intelligent life is not
an inevitable end product, and the fact of our existence places no
constraint on the probability (plif e(0), or plif e(moon)) of intelligent life
evolving on an earthlike planet. It might be close to 1, but it might just
as well be 10−30 (implying no other life within a Hubble radius). The
long intervals between critical events in the evolution of life on earth
suggest that evolution to intelligent life is unlikely to happen much
faster on other earthlike planets. The key requirements for the evolu-
tion of organic life on an earthlike planet are thus a source of organic
precursors, solid surfaces where the precursors can condense, long-
term maintenance of temperature within a range suitable for organic
reactions, and protection from hazards (e.g. impacts).
4. Possible influence of the moon on the evolution of life
Given on the one hand the wide-ranging consequences (compositional,
gravitational) of the earth having a large moon, and on the other the
stringent requirements for the origin and evolution of life, it would
perhaps be surprising if the former had not significantly affected the
latter. A number of specific suggestions have been made:
(1) Stabilisation of the earth's obliquity
Small changes in the earth's orbit and orientation probably drive cli-
moon3.tex; 2/06/2018; 12:17; p.3
4
C.R. Benn
matic change (Milankovitch theory, Imbrie 1982), and small changes
of the earth's obliquity (angle of spin axis with the perpendicular to
the orbital plane) of ∼ 1o could have triggered recent ice ages. Several
authors (e.g. Goldsmith & Owen 1980, Verschuur 1989) have noted
that a large moon would benefit life by stabilising the earth's obliquity,
and thus climate, and Laskar, Joutel & Robutel (1993) and de Surgy
& Laskar (1997) confirmed that in the absence of the moon, large
and chaotic variations of obliquity would have occurred. Mercury and
Venus have been stabilised by tidal dissipation (they spin very slowly),
but Mars, which has no large moon, undergoes chaotic variations of
obliquity in the range 0 - 60o (Laskar & Robutel 1993). Stabilisation of
the earth's obliquity might not have been crucial for the origin of life,
but could have been for the evolution of life on land.
(2) Elimination of the primordial atmosphere
Cameron & Benz (1991) and Taylor (1992) pointed out that the giant
impact which created the earth-moon system would have stripped the
earth of its thick primordial atmosphere, which might otherwise have
developed as has that of Venus, rendering the surface of the planet too
hot for organic life. On the other hand, the earth's atmosphere may
be thinner simply because most of the CO2 is locked up in carbonate
rocks.
(3) Generation of the earth's magnetic field
The earth's magnetic field partially shields the molecules of life from the
destructive effects of cosmic rays (although this may not be important
for submarine life). Compared to other solar-system bodies, the earth's
magnetic field is unusually strong for its angular momentum (though
the mechanisms are probably different for different bodies). Pearson
(1988) suggested that this anomaly might be due to the prolonged
heating of the earth's core following the impact that created the earth-
moon system.
(4) Generation of large tides
As noted above, a common theme in speculations about the origin of
the first self-replicating system is the importance of concentrating the
weak solution of organic molecules in the primordial sea, to encourage
polymerisation. The possible role of tidal pools, which repeatedly dry
out under the sun, has been stressed by many authors. Although the
amplitude of the tides raised by the moon is currently not much larger
than that of the tides raised by the sun, the moon was probably much
closer to the earth at the time of the origin of life (Chyba 1990), and
the tides raised would have been correspondingly larger, allowing tidal
pools with a much larger total area to be subjected to wetting/drying
cycles (e.g. Verschuur 1989, Gribbin & Rees 1990).
moon3.tex; 2/06/2018; 12:17; p.4
The moon and the origin of life
5
(5) Generation of longer-period tides
The length of time allowed for intertidal pools to dry out under the
sun before being refilled may also have been important. With longer
wetting/drying cycles, the probability of long sequences of chemical
reactions taking place is increased (the energies involved in organic
reactions are small, and they proceed slowly). It's possible that at the
time of the origin of life, the moon was not actually much closer to
earth than it is at present (Williams 1989, Taylor 1992). One may
then speculate (Rood & Trefil 1981) that beating between lunar and
solar tides to give neap/high tides at a longer interval (as observed at
present) was important, although of course longer intervals could also
be achieved by other means e.g. through seasonal effects. The hypoth-
esis has an interesting corollary. The condition for such beating is that
the strengths of the tides raised by the sun and by the moon, which
are ∝ density ∗ (angular diameter)3, are similar. The mean densities of
planetary bodies and main-sequence stars both happen to be ∼ atomic
(Carr & Rees 1979), so angular diameter ∝ (strength of tide)1/3 i.e.
the condition for long-period tides happens to imply similar angular
diameters of the sun and moon, as observed.1
With our current level of understanding of the origin and evolution
of life, the above five hypotheses remain speculative. However, the vari-
ety of suggested mechanisms (and more than one could be important)
attests to the far-reaching consequences of the earth having a large
moon. Some of these consequences inevitably impinge on events critical
to the development of life on earth. Thus it cannot be assumed that
this unusual feature of our environment is not anthropically selected.
5. Conclusions
In studies of the solar system, as in cosmology, we are dealing with
a unique example (so far), and we must beware the effects on our
observations of anthropic selection. The possibility that the presence
of the moon has affected the origin or evolution of life through one of
the mechanisms noted above implies that:
(1) hypotheses about the origin of the moon cannot be judged solely
on the basis of a priori likelihood (they need only satisfy equation 2);
(2) large moons may be useful pointers in the search for life-bearing
planets.
1 Curiously, this has a literary antecedent; in Martin Amis' novel London Fields
(Amis 1989) appears the line 'Perhaps that was the necessary condition of planetary
life: your sun must fit your moon'.
moon3.tex; 2/06/2018; 12:17; p.5
6
C.R. Benn
More generally, caution must be exercised in interpreting, and gen-
eralising from, unusual features of our local astrophysical environment;
they may turn out to be anthropically selected, and atypical. Indeed,
the earth and its environment may be very special (Ward & Brown-
lee 2000), and this might explain the puzzling lack of evidence for
intelligent life elsewhere in the universe (Tipler 1980, Wesson 1990).
References
Amis M., 1989, London Fields (Jonathon Cape), chapter 22
Barrow J.D., Tipler F.J., 1986, The Anthropic Cosmological Principle (Oxford
University Press)
Butler R.N., 1980, Irish Astr. J., 14, 516
Cameron A.G.W. & Benz W., 1991, Icarus, 92, 204
Carr B.J., Rees M.J., 1979, Nature, 278, 605, fig. 1
Chyba C.F., 1990, Nature, 343, 129
Comins N.F., 1993, What if the moon didn't exist? (Harper Collins)
Darwin C., 1871, in a letter to a friend
Deamer D.W., Barchfeld G.L., 1982, J. Mol. Evol., 18, 203
Deamer D.W., Fleischaker G.R., 1994, Origins of Life (Jones & Bartlett)
de Surgy O.N., Laskar J., 1997, Astron. Astroph., 318, 975
Efstathiou G., 1995, MNRAS, 274, L73
Goldsmith D., Owen T., 1980, The Search for Life in the Universe (Ben-
jamin/Cummings)
Gonzalez G., 1999, Astronomy & Geophysics, 40, issue 3, 18
Gribbin J., Rees M.J., 1990, Cosmic coincidences (Heinemann), chapter 11
Hartmann W.K., Phillips R.J., Taylor G.J., 1985, Origin of the Moon (Lunar and
Plan. Inst., Houston)
Imbrie J., 1982, Icarus, 50, 408
Kasting J.F., Whitmire D.P., Reynolds R.T., 1993, Icarus, 101, 108
Laskar J., Joutel F., Robutel P., 1993, Nature, 361, 615
Laskar J., Robutel P., 1993, Nature, 361, 608
Lazcano-Araujo A., Or´o J., 1981, Comets and the origin of life, ed. C. Ponnampe-
ruma (Reidel)
Lissauer J., 1997, Nature, 389, 327
Mishurov Yu. N., Zenina I.A., 1999, Astron. Astroph., 341, 82
Or´o J., Miller S.L., Lazcano A., 1990, Ann. Rev. Earth Planetary Sci., 18, 317
Pearson J., 1988, New Scientist, 25 August 1988, p.38
Ringwood A.E., 1990, in Origin of the Moon, eds. H.E. Newsom & J.H. Jones
(Oxford University Press)
Rood R.T., Trefil J.S., 1981, Are we alone? (Scribner)
Taylor S.R., 1992, Solar System Evolution (Cambridge University Press)
Tegmark M., Rees M.J., 1998, ApJ, 499, 526
Tipler F.J., 1980, Q. Jl. R. astr. Soc., 21, 267
Verschuur G.L., 1989, Sky & Telescope Nov 1989, p.452
Ward P.D., Brownlee D., 2000, Why complex life is uncommon in the Universe
(Copernicus)
Wesson P., 1990, QJRAS, 31, 161
moon3.tex; 2/06/2018; 12:17; p.6
The moon and the origin of life
7
Wetherill G., 1995, Nature, 373, 470
Whittet D., 1997, Astronomy & Geophysics, 38, issue 5, p.8
Williams G.E., 1989, J.Geol.Soc.London, 146, 97
moon3.tex; 2/06/2018; 12:17; p.7
moon3.tex; 2/06/2018; 12:17; p.8
|
astro-ph/9909470 | 1 | 9909 | 1999-09-28T17:23:05 | Radiation-Pressure Ejection of Planetary Nebulae in Asymptotic-Giant- Branch Stars | [
"astro-ph"
] | We have investigated the possibility that radiation pressure might trigger planetary nebula (PN) ejection during helium-shell flashes in asymptotic- giant-branch (AGB) stars. We find that the outward flux at the base of the hydrogen envelope during a flash will reach the Eddington limit when the envelope mass falls below a critical value that depends on the core mass and composition. These results may help to explain the helium-burning PN nuclei found in the Magellanic Clouds. | astro-ph | astro-ph |
Radiation-Pressure Ejection of Planetary Nebulae in
Asymptotic-Giant-Branch Stars
A. V. Sweigart
NASA Goddard Space Flight Center, Code 681, Greenbelt, MD 20771
Abstract. We have investigated the possibility that radiation pressure
might trigger planetary nebula (PN) ejection during helium-shell flashes
in asymptotic-giant-branch (AGB) stars. We find that the outward flux
at the base of the hydrogen envelope during a flash will reach the Edding-
ton limit when the envelope mass Menv falls below a critical value that
depends on the core mass MH and composition. These results may help
to explain the helium-burning PN nuclei found in the Magellanic Clouds.
1. Description of Radiation-Pressure Instability
We have computed extensive AGB evolutionary sequences for heavy-element
abundances Z = 0.01716 (solar) and 0.002 in order to study a radiation-pressure
instability for ejecting a PN during a helium-shell flash. The existence of such
an instability was previously confirmed by Wood & Faulkner (1986) but only for
more massive AGB stars with large core masses (MH > 0.86 M⊙). In contrast,
we find this instability at core masses as small as MH ≈ 0.6 M⊙ due, in part,
to our use of the new OPAL opacities (Rogers & Iglesias 1992).
Results for a typical sequence are presented in the left panels of Figure 1.
The strong flashes in this sequence cause the hydrogen shell to expand outward
to very low temperatures and densities. This, in turn, lowers the gas pressure
and hence the minimum value βmin of the ratio β of gas to total pressure near
the base of the hydrogen envelope. Most importantly, we note from Figure 1
that βmin decreases monotonically from flash to flash and eventually goes to 0,
implying that the entire envelope is then supported by radiation pressure.
During the final flashes in Figure 1 the temperature within the hydrogen
shell falls to log T = 5.3 at which there is a well-known peak in the OPAL
opacities. When the shell reaches this peak, its opacity suddenly increases, and,
as a result, its Eddington luminosity abruptly drops. This occurs at a time
when the outward flux within the shell is considerably enhanced by the flash.
As the shell continues to cool, its opacity increases further, thereby forcing an
even greater expansion and driving β down to even smaller values. The star thus
encounters an "opacity catastrophe" whose likely outcome is envelope ejection.
The present sequences show that there is a critical Menv at which radiation
pressure will support the hydrogen envelope during a flash (see right panel of
Figure 1). This critical Menv increases linearly with MH and decreases with
decreasing metallicity, since the peak in the OPAL opacities is then smaller.
The critical values of Menv in Figure 1 represent the minimum envelope mass at
1
Figure 1. Helium-burning luminosity LHe in solar units (upper left panel)
and minimum value βmin of the ratio β of gas to total pressure near the
base of the hydrogen envelope (lower left panel) as a function of the
core mass MH during the helium-shell flashes of a solar metallicity AGB
star. The right panel shows the dependence of the critical envelope mass
Menv at which β goes to 0 during a flash on MH for two heavy-element
abundances: Z = 0.01716 (solid squares) and 0.002 (open squares). The
dashed lines are linear fits to the model data.
which an AGB star can undergo a flash without β going to 0. These results
depend, however, on the extent of 3rd dredge-up in the model calculations.
The minimum in β during a flash occurs at the time when 3rd dredge-up
is most likely. We suggest that such dredge-up might trigger PN ejection by
increasing the opacity and thereby lowering the Eddington luminosity at the
base of the envelope, especially at low metallicites where dredge-up is favored
by current theoretical models. This might explain why the majority of halo PN's
and most non-Type I PN's in the Magellanic Clouds are carbon rich. It might
also explain why the PN K648 in M15 has a high carbon abundance despite the
lack of AGB carbon stars in this globular cluster. We suggest therefore that the
high carbon abundance of K648 was the cause of its PN ejection.
References
Rogers, F. J., & Iglesias, C. A. 1992, ApJS, 79, 507
Wood, P. R., & Faulkner, D. J. 1986, ApJ, 307, 659
2
|
astro-ph/0306096 | 1 | 0306 | 2003-06-05T18:58:37 | Self-Similar Models for the Mass Profiles of Early-type Lens Galaxies | [
"astro-ph"
] | We introduce a self-similar mass model for early-type galaxies, and constrain it using the aperture mass-radius relations determined from the geometries of 22 gravitational lenses. The model consists of two components: a concentrated component which traces the light distribution, and a more extended power-law component (rho propto r^-n) which represents the dark matter. We find that lens galaxies have total mass profiles which are nearly isothermal, or slightly steeper, on the several-kiloparsec radial scale spanned by the lensed images. In the limit of a single-component, power-law radial profile, the model implies n=2.07+/-0.13, consistent with isothermal (n=2). Models in which mass traces light are excluded at >99 percent confidence. An n=1 cusp (such as the Navarro-Frenk-White profile) requires a projected dark matter mass fraction of f_cdm = 0.22+/-0.10 inside 2 effective radii. These are the best statistical constraints yet obtained on the mass profiles of lenses, and provide clear evidence for a small but non-zero dark matter mass fraction in the inner regions of early-type galaxies. In addition, we derive the first strong lensing constraint on the relation between stellar mass-to-light ratio (Upsilon) and galaxy luminosity (L): Upsilon propto L^[0.14 (+0.16)(-0.12)], which is consistent with the relation suggested by the fundamental plane. Finally, we apply our self-similar mass models to current problems regarding the interpretation of time delays and flux ratio anomalies in gravitational lens systems. | astro-ph | astro-ph | Self-Similar Models for the Mass Profiles of Early-type Lens Galaxies
D. Rusin1, C.S. Kochanek1, C.R. Keeton2,3
ABSTRACT
We introduce a self-similar mass model for early-type galaxies, and constrain it using
the aperture mass-radius relations determined from the geometries of 22 gravitational
lenses. The model consists of two components: a concentrated component which traces
the light distribution, and a more extended power-law component (ρ ∝ r−n) which
represents the dark matter. We find that lens galaxies have total mass profiles which
are nearly isothermal, or slightly steeper, on the several-kiloparsec radial scale spanned
by the lensed images.
In the limit of a single-component, power-law radial profile,
the model implies n = 2.07 ± 0.13, consistent with isothermal (n = 2). Models in
which mass traces light are excluded at > 99% confidence. An n = 1 cusp (such as
the Navarro-Frenk-White profile) requires a projected dark matter mass fraction of
fcdm = 0.22 ± 0.10 inside 2 effective radii. These are the best statistical constraints yet
obtained on the mass profiles of lenses, and provide clear evidence for a small but non-
zero dark matter mass fraction in the inner regions of early-type galaxies. In addition,
we derive the first strong lensing constraint on the relation between stellar mass-to-light
−0.12 , which is consistent with the relation
suggested by the fundamental plane. Finally, we apply our self-similar mass models to
current problems regarding the interpretation of time delays and flux ratio anomalies
in gravitational lens systems.
ratio Υ and galaxy luminosity L: Υ ∝ L0.14+0.16
Subject headings: galaxies: elliptical and lenticular -- galaxies: structure -- gravitational
lensing
3
0
0
2
n
u
J
5
1
v
6
9
0
6
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
The structure of galaxies is closely linked to their formation and evolution, and therefore
provides a vital testing ground for the cold dark matter (CDM) paradigm which has proven so
successful on large scales (e.g., Spergel et al. 2003). The primary observable is the shape of the
radial mass profile, from which the relative distributions of luminous and dark matter can be
determined. The outermost regions of galaxies have now been extensively probed using weak lensing
1Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138
2Astronomy and Astrophysics Department, University of Chicago, 5640 S. Ellis Ave., Chicago, IL 60637
3Hubble Fellow
-- 2 --
(e.g., Fischer et al. 2000; McKay et al. 2001; Kleinheinrich et al. 2003) and satellite dynamics (e.g.,
Zaritsky et al. 1997; Romanowsky & Kochanek 2001; McKay et al. 2002; Prada et al. 2003), each
of which reveals an extended dark matter halo out to 100 − 300 kpc. Methods for tracing the inner
regions of galaxies largely depend on the galaxy morphology. Spiral galaxies are the most easily
studied, and are consequently the best understood, as rotation curves can be mapped to ∼ 30 kpc
using kinematic tracers such as HI. Flat rotation curves at & 5 kpc clearly indicate the presence of
dark matter (e.g., Rubin, Thonnard & Ford 1980), but dynamical observations also suggest that
the inner halo may be significantly less concentrated (e.g., McGaugh & de Blok 1998; de Blok et al.
2001; Salucci 2001; de Blok & Bosma 2002) than the cuspy halos predicted by CDM simulations
(e.g., Moore et al. 1999b; Bullock et al. 2001).
Compared to their late-type counterparts, early-type (elliptical and lenticular; E/S0) galaxies
are not as well understood. Some of the best evidence for dark matter outside of a few optical radii
has been obtained from investigations of the extended, X-ray emitting gas (e.g., Fabbiano 1989;
Matsushita et al. 1998; Loewenstein & White 1999). The inner regions of early-type galaxies are
typically probed using stellar dynamics (e.g., Rix et al. 1997; Gerhard et al. 2001). However, the
velocity dispersion is not an unambiguous tracer of the mass profile, as it requires an understanding
of the orbital anisotropy. In addition, the existence of a tight fundamental plane (FP; Djorgovski
& Davis 1987; Dressler et al. 1987) presumably provides clues about the inner structure of early-
type galaxies, but little consensus has been achieved regarding its meaning (e.g., Faber et al. 1987;
Renzini & Ciotti 1993; van Albada, Bertin & Stiavelli 1995; Ciotti, Lanzoni & Renzini 1996; Pahre,
de Carvalho, & Djorgovski 1998; Bertin, Ciotti, & Del Principe 2002). Many outstanding questions
therefore remain regarding the structure of E/S0 galaxies, the universality of this structure, and
its relation to the predictions of CDM models.
Strong gravitational lensing is a powerful and unique tool for investigating the mass distri-
butions in early-type galaxies at intermediate redshift, as it probes mass directly. First, model-
independent projected masses may be inferred from the geometry of the lensed images (e.g., Schnei-
der, Ehlers & Falco 1992). When mass models are normalized by this constraint, the stellar velocity
dispersion becomes a far more sensitive probe of the radial mass profile. Ongoing efforts by the
Lens Structure and Dynamics (LSD) survey (Koopmans & Treu 2002, 2003; Treu & Koopmans
2002a; collectively "KT") have made significant progress in obtaining and interpreting dynami-
cal measurements in gravitational lens systems. Second, models of the galaxy mass distribution
can be constructed based on the positions, morphologies and flux ratios of the lensed images. Of
particular importance are the additional astrometric constraints obtained from complex structure
in the lensed source, which can be used to break degeneracies and distinguish among radial mass
profiles. These constraints include large-scale extended emission from radio lobes (e.g., Kochanek
1995) or quasar host galaxies (see, e.g., Kochanek, Keeton, & McLeod 2001 and Saha & Williams
2001 for general discussions), lensed images of multi-component sources (e.g., Cohn et al. 2001;
Munoz, Kochanek & Keeton 2001), and the relative orientations of milliarcsecond-scale radio jets
(e.g., Rusin et al. 2002; Winn, Rusin & Kochanek 2003).
-- 3 --
The mass distributions in several lens galaxies have now been investigated using either direct
modeling or stellar dynamics.
In all but one case, the profile is consistent with an isothermal
model (ρ ∝ r−2). The lone exception is PG1115+080, for which the high stellar velocity dispersion
(Tonry 1998) suggests a mass profile steeper than isothermal (ρ ∝ r−2.35; Treu & Koopmans
2002b), although this also implies that the galaxy lies well off the FP. The lensing results are
therefore generally consistent with the nearly isothermal profiles favored by X-ray and dynamical
studies of local elliptical galaxies.
Understanding the mass profiles of lens galaxies is required for the determination of the Hubble
constant (H0 ≡ 100h km s−1 Mpc−1) from measured time delays. Kochanek (2002) demonstrates
that the predicted delay between two images, and hence the derived value of H0, is primarily
governed by the average surface mass density in the annulus defined by their radial distances from
the lens center, with a small correction for the slope of the profile in this annulus. Each of these
quantities is determined by the shape of the radial mass profile. Based on a study of several lens
systems which are dominated by a single galaxy and have well-measured time delays, Kochanek
(2002) finds h = 0.51 ± 0.05 if lenses are isothermal. This value is lower than those obtained
from the Hubble Space Telescope (HST) Key Project (h = 0.72 ± 0.08; Freedman et al. 2001) and
WMAP data (h = 0.72± 0.05, although this requires the assumption of a flat cosmology and a dark
energy equation of state of w = −1; Spergel et al. 2003). In fact, Kochanek (2002, 2003) shows
that early-type galaxies must be described by models in which mass traces light, and hence have
no extended dark matter halos, to reconcile current time delay measurements with h ≃ 0.7 (see
also Williams & Saha 2000). This apparent discrepancy with other observational constraints and
the CDM paradigm demands continued examination of the lens galaxy population.
Decomposing the luminous and dark matter components of early-type galaxies is also poten-
tially important for investigating the nature of flux ratio discrepancies in gravitational lens systems.
It has now been extensively demonstrated that simple, smooth mass models have great difficulty
reproducing the observed flux ratios in four-image lenses (e.g., Dalal & Kochanek 2002; Metcalf
& Zhao 2002; Chiba 2002; Keeton, Gaudi & Petters 2002), both in the optical and radio. The
effect has been traced to small-scale structure in the gravitational potential (Kochanek & Dalal
2003), and can be due to either stars (microlensing; e.g., Witt, Mao, & Schechter 1995; Schechter
& Wambsganss 2002) or CDM satellite halos (millilensing; e.g., Mao & Schneider 1998; Metcalf
& Madau 2001; Bradac et al. 2002; Chiba 2002; Dalal & Kochanek 2002). Both phenomena are
surely present, as radio sources are generally too large to be affected by stars, while uncorrelated
time variability is an unmistakable signature of stellar microlensing in some optical lenses (e.g.,
Wozniak et al. 2000). Microlensing is most efficient at accounting for observed anomalies, in partic-
ular the preferred demagnification of the saddle point image, when stars comprise a small fraction
(15 -- 30%) of the surface mass density at the image radius (Schechter & Wambsganss 2002). If we
can constrain a two-component model of the galaxy mass distribution, then we can predict the
stellar surface mass fraction at the radii typically spanned by lensed images. Thus, the predictions
and requirements of the microlensing hypothesis may be tested.
-- 4 --
Because mass distributions have been directly constrained in only a small number of gravita-
tional lenses, it is worthwhile to consider statistical constraints on the lens galaxy population. Such
attempts have thus far been very limited. Several analyses (e.g., Rusin & Ma 2001; Evans & Hunter
2002) have focused on the absence of central images in deep radio maps of gravitational lens sys-
tems, which has been invoked to rule out profiles that are much shallower than isothermal or have
significant cores. Keeton (2003) suggests that such constraints have been over-interpreted, since
the magnification of the missing image is strongly dependent on the properties of the mass profile
very close to the galaxy center. In addition, because mass distributions steeper than isothermal do
not produce central images, any constraints based on the absence of these images are one-sided.
The number and sizes of lens systems have also been used to investigate galaxy mass profiles in the
context of CDM models (Keeton 2001). It is shown that adiabatic compression (e.g., Blumenthal
et al. 1986) yields nearly flat rotation curves for a range of initial halo concentrations and cooled
baryon fractions (see also Kochanek & White 2001 and Kochanek 2003), but typical CDM halos
(e.g., Bullock et al. 2001) appear to be too centrally concentrated to account for the data. The
analysis, however, offers few quantitative results describing the mean mass profile and scatter, and
therefore demands further study of two-component mass models.
In this paper we introduce a new technique to statistically constrain the mass profiles of early-
type lens galaxies via aperture mass measurements. Each lens must obey a strict relationship
between the image radii and the projected masses they enclose. The set of mass-radius relations
can be used to investigate the typical mass profile in a sample of lenses if we assume that early-type
galaxies have a self-similar (homologous) structure. This means that the functional form (shape) of
the mass distribution is the same from galaxy to galaxy, and is scaled by the properties of the light
distribution. An underlying homology between mass and light is suggested by the existence of a
tight fundamental plane relating optical and dynamical observables. In one popular interpretation
of the FP (e.g., Faber et al. 1987; van Albada et al. 1995), early-type galaxies are structurally
homologous, but the mass-to-light ratio has a luminosity dependence. Gerhard et al. (2001) use
dynamical data to bolster this hypothesis, demonstrating that more luminous early-type galaxies
have higher stellar mass-to-light ratios in their central regions. Other analyses have cast doubt on
the description of the FP in terms of strong homology and varying mass-to-light ratios (e.g., Caon,
Capaccioli, & D'Onofrio 1993; Pahre et al. 1998; Bertin et al. 1994, 2002), and instead suggest
a weak homology in which structural parameters vary systematically with luminosity. However,
because a strong homology is the simplest reasonable assumption, it is a good starting point for
the statistical investigation of mass profiles in early-type galaxies.
Section 2 details the aperture mass-radius constraints offered by lensing. Section 3 describes
the lens sample and outlines our various assumptions. Section 4 introduces our self-similar models
and the associated methodology. Section 5 presents our statistical constraints on the radial mass
distribution in lenses. Section 6 demonstrates how these constraints can be used for the interpreta-
tion of measured time delays and flux ratio anomalies among lensed images. Section 7 discusses our
findings in the context of recent astrophysical and cosmological results. Unless otherwise stated,
we assume a flat ΩM = 0.3 cosmology with h = 0.65 for all calculations.
-- 5 --
2. Aperture Masses from Lens Geometries
The geometries of gravitational lens systems yield model-independent constraints on projected
masses. Consider, for example, a deflector with a circularly symmetric surface density distribution,
and a source which sits directly behind it. The image will form on a ring of (physical) Einstein
radius REin, which is related to the aperture mass (MEin) it encloses by
1
π
MEin
R2
Ein
=
c2
4πG
Ds
DdDds ≡ Σcr ,
(1)
where Σcr is the critical surface density. Angular diameter distances to the lens, to the source,
and from the lens to the source are Dd, Ds and Dds, respectively. For this simple case, the mass
enclosed by the ring is determined by its radius alone, and is therefore independent of the shape of
the mass profile.
Real lenses are more complicated. First, sources (and their images) are, in general, not sym-
metrically placed with respect to the mass distribution. This means that the geometry may be
described by more than one characteristic radius. Second, gravitational potentials are, in general,
not circularly symmetric. The presence of a quadrupole, due to either external shear or ellipticity,
effectively stretches the image plane along one axis and compresses it along a perpendicular axis,
thereby confusing the interpretation of geometry in terms of mass. These issues are surmountable,
however, allowing any lens geometry to yield model-independent relations between masses and
radii.
Four-image lenses (quads) offer the best mass constraints, as the above complications are
minimized. The images typically reside at similar radii, and their good angular coverage allows
for the robust determination and removal of the quadrupole. Just about any mass profile can
reproduce the image positions and fluxes in a quad, but modeling demonstrates that there is a
radius at which all models agree on the enclosed projected mass (e.g., Cohn et al. 2001; Munoz
et al. 2001). We have tested the generality of this claim using Monte Carlo simulations. First,
we create fake four-image lenses from an elliptical mass distribution with some fixed radial profile
residing in an external shear field of random orientation. We then model the image positions and
fluxes using a broad range of mass profiles, tabulating the mass enclosed by a circular aperture as
a function of radius for each best-fit model. We find that the suite of aperture mass curves M (R)
cross at nearly the same radius (REin), the value of which can be accurately estimated (to ∼ 1%)
from the critical radius of a model-fit singular isothermal sphere (SIS) in an external shear field.
This is true regardless of whether the mass distribution is spherical or elliptical. The enclosed,
projected mass inside REin is then given by eq. (1), to an accuracy of ∼ 2%.
Two-image lenses (doubles) yield weaker mass constraints. If the potential is circularly sym-
metric, then the lens equation dictates a model-independent mass-radius relation:
-- 6 --
1
π (cid:20) M (R1)
R1
+
M (R2)
R2
(cid:21)(cid:18)
1
R1 + R2(cid:19) = Σcr,
(2)
where R1 and R2 are the radii of the images with respect to the galaxy center, and M (R1) and
M (R2) are the projected masses they enclose. Hence, doubles constrain a combination of two radii
and two masses. Note that in the limit of a ring (R1 = R2), we recover eq. (1). The presence of
a quadrupole smears the mass-radius relation for doubles, and the limited number of constraints
means that its effects cannot be removed in a model-independent manner. If the quadrupole is
represented as an external shear, then the maximum fractional deviation from eq. (2) is equal to
the shear magnitude γ, and this occurs when the shear axis is parallel or perpendicular to the axis
defined by the lensed images. Averaging over orientations, the rms deviation is ≃ γ/√2. We have
confirmed this assertion using Monte Carlo simulations, in which fake doubles are produced from
one mass profile and modeled using a range of mass profiles. While four-image lenses typically
have quadrupoles which can be described by shear fields of amplitude 0.1 < γ < 0.2 (e.g., Keeton,
Kochanek & Seljak 1997; Holder & Schechter 2003), we might expect doubles to have somewhat
lower quadrupoles, on average, as they represent a greater fraction of the lensing cross section
when the shear is small (e.g., Keeton et al. 1997; Rusin & Tegmark 2001; Finch et al. 2002).
Consequently, deviations from eq. (2) should generally be ∼ 10%. We therefore employ eq. (2)
as a model-independent mass-radius relation for doubles, and set a 10% tolerance for quadrupole
effects.4
3. Data
3.1. Lens Sample
We begin with the sample of 28 early-type gravitational lens galaxies and bulges analyzed
by Rusin et al. (2003). Note that this includes Q2237+050 (Huchra et al. 1985), which is lensed
entirely by the bulge of a spiral galaxy. Twenty-two of the systems have measured lens redshifts
(zd), and 18 of these are in combination with a measured source redshift (zs).5 Previously, for
systems with no zd, the lens redshift had been estimated under the requirement that the galaxy fall
on the fundamental plane (Kochanek et al. 2000). To make use of this technique, one must first
4We note in passing that when using steep radial profiles to fit fake lenses produced from shallower profiles, huge
external shears or extremely flattened deflectors may be required. These solutions, however, can be rejected on the
grounds of physical implausibility. First, the mass distributions of lens galaxies would have to be significantly more
flattened than the light distributions, which typically have moderate axial ratios of ∼ 0.7 − 0.8; (e.g., Keeton et al.
1997). Second, such large external shears would have to be directly traced to nearby, cluster-scale perturbers, which
are not observed in the relatively low-density environments of lens galaxies (Keeton, Christlein & Zabludoff 2000).
5Here we include a recent measurement of zs = 1.17 (Treu & Koopmans 2003) for MG1549+3047 (Leh´ar et al.
1993).
-- 7 --
estimate the stellar velocity dispersion, which necessitates an explicit assumption of a galaxy mass
profile. Since this paper is concerned with constraining mass models, it is prudent to exclude the
six lenses without spectroscopic lens redshifts. The number of lenses used in the following analysis
is therefore 22. The systems are listed in Table 1.
3.2. Geometric Properties
The important geometric properties of each lens system are derived from the radii of the
lensed images. Image positions with respect to the lens galaxy are determined from the HST data,
based on the fitting methods outlined by Leh´ar et al. (2000).6 The uncertainties on these positions
are typically . 5 milliarcseconds, which is negligible compared to the radii of even the closest
lensed images in doubles (∼ 0.1 arcsec). For quads and rings, the angular Einstein radius rEin is
determined by fitting an SIS plus external shear model to the lens data, and then converted to the
physical radius REin = rEinDd. The Einstein radius can typically be derived to ∼1% precision, so
its uncertainty is also negligible. The geometric properties of the lens sample are listed in Table 1.
Interpreting radii in terms of masses requires the critical surface density Σcr, which depends on
source and galaxy redshifts through the angular diameter distances. Recall that each lens galaxy
in our sample has a measured redshift. If the source redshift has also been measured, then there
is no uncertainty in Σcr, at least within the context of our standard cosmological assumptions. If
the source redshift is not known, we derive the uncertainty in Σcr (actually log Σcr) with Monte
Carlo techniques. Following Rusin et al. (2003), we draw 10000 values from a Gaussian distribution
zs = 2.0±1.0 for the source redshift, keeping only trials with 1 ≤ zs ≤ 5 because virtually all known
lensed sources lie within this range. We adopt the zs = 2.0 value of log Σcr as the median, and the
rms scatter around this value as the uncertainty in log Σcr.
3.3. Photometric Properties
The important photometric properties of each galaxy are the intermediate axis effective radius
and luminosity. These quantities are derived from HST data obtained with the WFPC2 and
NICMOS cameras. The fitting methods are detailed by Leh´ar et al. (2000), Kochanek et al. (2000)
and Rusin et al. (2003). The optical effective radius (re; or Re = reDd in physical units) is
determined by fitting a de Vaucouleurs profile to the galaxy in the filter with the highest signal-to-
6Because the two images of the radio lens B2319+051 (Rusin et al. 2001) have not been detected in the optical, a
direct determination of their positions with respect to the galaxy is currently impossible. However, the structure in
the lensed jets allows the galaxy position to be determined from mass modeling. Rusin et al. (2001) demonstrate this
in the context of an isothermal mass profile, but our calculations show that the recovered galaxy position is nearly
independent of the assumed model. We therefore include this system in our analysis, using the estimated R1 and R2
from that paper.
-- 8 --
noise ratio. This quantity is then held fixed to determine the mean surface brightness within the
effective radius (µe,Y ) in all filters Y . The total magnitude is then mY = µe,Y − 5 log re − 2.5 log 2π.
The photometric properties of our sample are tabulated in Rusin et al. (2003).
Complications arise from the need to convert the observed galaxy magnitudes to some standard
luminosity scale that accounts for K-corrections and luminosity evolution. As we shall see in §4,
only the magnitude offset M − M∗ = −2.5 log(L/L∗) is required for our analysis, since we can
parameterize the homology model in terms of the present-day (z = 0) stellar mass-to-light ratio for
an L∗ galaxy. Assuming that M − M∗ does not evolve (i.e., that the evolution rate is independent
of luminosity; the same assumption was made in Rusin et al. 2003 and all other analyses using
the FP to study galaxy evolution; e.g., van Dokkum et al. 1998, 2001; Treu et al. 2001, 2002), we
can estimate this quantity from the observed magnitudes, without converting to rest frame bands.
Specifically, we compare the observed magnitudes (mobs,Y ) to models (mmod,Y ) for the apparent
magnitudes of an L∗ galaxy at the lens redshift, and then calculate the mean offset:
hM − M∗i = PY [mobs,Y − mmod,Y ]/(δmobs,Y )2
PY 1/(δmobs,Y )2
.
(3)
We compute the model magnitudes by convolving spectral energy distributions from the GISSEL96
version of the Bruzual & Charlot (1993) spectral evolution models with transmission curves for HST
filters (available from the technical archives of STScI, with zero-points from Holtzman et al. 1995).
All models are normalized to a fixed present-day characteristic magnitude of M∗ = −19.9 + 5 log h
for early-type galaxies in the B band (e.g., Madgwick et al. 2002). We take as a fiducial model
an instantaneous starburst at zf = 3 with solar metallicity Z = Z⊙ and a Salpeter (1955) initial
mass function (IMF). The assumed metallicity is consistent with observations of early-type field
galaxies at 0.3 < z < 0.9 (Ferreras, Charlot & Silk 1999), and the assumed formation redshift
is consistent with a number of analyses which strongly favor old stellar populations (mean star
formation redshift hzfi > 2; e.g., Bernardi et al. 1998; Schade et al. 1999; Kochanek et al. 2000;
van Dokkum et al. 2001; Im et al. 2002; van de Ven, van Dokkum & Franx 2002; Rusin et al. 2003).
Considering a broad range of stellar models (1.5 < zf < 5, 0.4 < Z/Z⊙ < 2.5) yields an rms scatter
in hM − M∗i of 0.1 -- 0.2 mag for most lenses. This uncertainty is much larger than the measurement
errors, since all of the 22 galaxies have excellent photometry. We therefore adopt a luminosity error
of δ(M − M∗) = 0.20, or δ log(L/L∗) = 0.08, for all lens galaxies.
Estimating L/L∗ requires some understanding of luminosity evolution, which is implicitly
included in the spectral models. For example, when normalized to a fixed L∗ at z = 0, a faster
evolution rate (lower hzfi) would predict higher values of L∗(z), and hence we would derive lower
values of L/L∗ for each lens galaxy. If the true mean star formation redshift is hzfi < 1.5, then
our fiducial model of hzfi = 3 could greatly overestimate L/L∗, especially for high redshift lenses
(zd > 0.7). However, most studies of early-type field galaxies are inconsistent with the rapid
luminosity evolution implied by hzfi < 1.5. We note in passing that if the evolution rate is
explicitly fit by our mass models (§4), the scatter is minimized by hzfi ≃ 3, and increases for
later formation redshifts. Unfortunately, the resulting constraints on the evolution rate are very
-- 9 --
poor, as a number of other parameters are being fit. Since this paper focuses on the structure of
lens galaxies, we believe it is best to simply assume a reasonable mean star formation redshift of
hzfi = 3. We reiterate that our estimated uncertainties on L/L∗ are sufficient to account for the
spread in evolution rates over the favored range of hzfi > 1.5.
Finally, note that many of the quantities entering our analysis are correlated. For example, the
total magnitude and the fraction of the light enclosed by the aperture radius R both correlate with
the effective radius Re. Neither effect is very significant, because δ log Re is typically quite small
(see Rusin et al. 2003), but for completeness we take these correlations into account via Monte
Carlo calculations. We draw 10000 trials from a Gaussian distribution representing the effective
radius (mean log Re, width δ log Re), compute the derived quantities, and store them for later use.
We use the 10000 sets of correlated parameters, plus the independent (uncorrelated) luminosity
error of δ log(L/L∗) = 0.08, to calculate the scatter in the quantities entering the fitting functions
given below.
4. Self-similar Models for Galaxy Mass Distributions
The aperture mass-radius relation (eq. 1 or 2) for a single lens tells us nothing about the mass
profile. However, recall that the image splitting scale depends on the redshifts of both the lens
galaxy (zd) and lensed source (zs) through the angular diameter distances in Σcr. For example,
consider a spherical galaxy with a power-law mass density ρ(r) = ρ0(r/r0)−n, and a corresponding
surface density Σ(R) = Σ0(R/R0)1−n. If a source sits directly behind this galaxy, it will be lensed
into a ring of Einstein radius REin = [2Σ0/Σcr(3−n)]1/(n−1)R0. We could therefore imagine placing
a fixed mass distribution at various zd, where it will lens sources at various zs. The resulting set
of Einstein radii REin and aperture masses MEin would then allow us to trace out the radial mass
profile of that galaxy.
Nature offers us an ensemble of gravitational lens systems with a wide range of lens (0 . zd . 1)
and source (1 . zs . 5) redshifts, but the galaxy properties are far from uniform. The aperture
mass-radius relations can, however, determine the typical (or "mean") mass profile of the lens galaxy
population if all galaxies can be placed on a common mass scale. This is most easily accomplished by
postulating some self-similar (homology) model to relate the mass and light distributions of early-
type galaxies. A given model allows projected masses M (R) to be predicted for each lens using the
photometric parameters, which can then be compared to the mass-radius relations determined from
the lensing geometries. The large range of image radii in our sample, spanning 0.2 . R/Re . 7,
gives us a long baseline for mapping the profile. We note, however, that the distribution is not
uniformly populated, with most of the quads and rings probing the range 1 . REin/Re . 4.
Doubles help fill in the radial coverage: the inner images typically span 0.4 . R1/Re < 1, while
the outer images span 2 . R2/Re . 4.
We consider a physically-motivated two component (luminous plus dark matter) model for the
-- 10 --
mass distribution. Because of the limited sample of lenses on which to test the model, we keep
the number of parameters to a minimum. The component which traces the luminous matter is
modeled by two parameters: the mean present-day (z = 0) stellar mass-to-light ratio Υ∗ (in the
rest frame B band) for an L∗ galaxy, and an exponent (x) describing its dependence on the galaxy
luminosity L: Υ ∝ Lx. Under the assumption of homology, the fundamental plane implies that
x ≃ 0.3 (e.g., Jorgensen, Franx & Kjaergaard 1996). For each galaxy, the luminous surface mass
density is represented by a de Vaucouleurs profile, and the associated mass enclosed by an aperture
R is
g(cid:18) R
Re(cid:19) ,
(4)
Mlum(R) = Υ∗ L∗ (cid:18) L
L∗(cid:19)1+x
where g(R/Re) is the projected fraction of the luminosity inside R. Recall that for the de Vau-
couleurs profile, g(0) = 0, g(1) = 0.5 and g(∞) = 1. Note that our use of the intermediate axis
effective radius, which is the geometric mean of the major and minor axes, allows us to accurately
determine the fraction of the luminosity enclosed by a circular aperture, even when the surface
brightness distribution is elliptical.
We model the dark matter as a power-law mass distribution. N-body simulations predict halos
in which the mass density follows a shallow power-law slope at small radius, and turns over to a
steeper slope beyond some break radius (e.g., Navarro, Frenk & White 1997, hereafter NFW; Moore
et al. 1999b). However, because the lensed image radii probe only the inner several kiloparsecs of the
mass distribution, we believe that a power-law approximation for the dark matter halo is reasonable
on this scale. We describe this component by two parameters: the projected mass fraction (fcdm)
within 2Re, which is roughly the median scale of the Einstein radii of our lens sample, and the
logarithmic slope of the mass density profile (n, where ρ ∝ r−n).7 For each galaxy, the dark matter
enclosed by an aperture R is
Mcdm(R) = Mlum(2Re)
fcdm
1 − fcdm (cid:18) R
2Re(cid:19)3−n
= Υ∗ L∗(cid:18) L
L∗(cid:19)1+x
g(2)
fcdm
1 − fcdm (cid:18) R
2Re(cid:19)3−n
where g(2) = 0.69. The total model-predicted mass inside R is then
Mmod(R) = Mlum(R) + Mcdm(R) .
, (5)
(6)
Note that the parameters Υ∗ and x set the normalization of the mass profile, while the dark matter
parameters fcdm and n modulate its shape. The homology scheme effectively places all galaxies on
a common mass scale by normalizing the aperture masses M (R) by (L/L∗)1+x, and the aperture
radii R by Re. If a strict homology holds, then all galaxy aperture masses and radii should reside
on a curve in the space of M (R)/(L/L∗)1+x versus R/Re.
7The parameter n should be thought of as a "local" density slope.
In reality, our fits use a surface density
Σ ∝ R1−n. This avoids the divergences encountered in converting between volume and surface densities for n ≤ 1.
-- 11 --
Modeling, dynamical and statistical studies of gravitational lens galaxies often make use of a
single-component mass distribution with a scale-free radial profile. While simplistic, this model al-
lows one to explore much of the lensing phenomenology related to mass concentration, and provides
a standard for comparing profile constraints from different lens systems. It is therefore worthwhile
to consider the scale-free limit of our homology model. By taking Υ∗ → 0 and fcdm → 1 (such that
Υ∗fcdm/(1 − fcdm) → const.), we obtain a pure power-law mass distribution:
Mpl(R) = M0(cid:18) L
L∗(cid:19)1+x(cid:18) R
2Re(cid:19)3−n
,
(7)
where M0 is the projected mass inside 2Re for an L∗ galaxy.
We will constrain the parameters of the homology model by the set of mass-radius relations
from the lens sample. The optimization is performed using the goodness-of-fit function
χ2 = χ2
qr + χ2
d .
(8)
In quads and rings, the lensing geometry constrains the mass enclosed by the Einstein radius (eq. 1).
These systems are evaluated using
χ2
qr =Xi " log Mmod(REin,i) − log[Σcr,iπR2
δscaleδi
Ein,i]
#2
,
(9)
where δscale and δi are defined below. In doubles, the lensing geometry constrains a combination
of two radii and projected masses (eq. 2). These systems are evaluated using
χ2
d =Xi (cid:20) log[Mmod(R1,i)/R1,i + Mmod(R2,i)/R2,i] − log[Σcr,iπ(R1,i + R2,i)]
δscaleδi
(cid:21)2
.
(10)
The logarithmic uncertainty δi on each data point is derived using the Monte Carlo methods
outlined in §3, but is well approximated as δ2
γ, where
δγ is the additional 10% tolerance for quadrupole-related smearing of the mass-radius relation in
doubles. We set δγ = 0 for quads and rings, and δ log Σcr,i = 0 for systems with a measured zs.
Finally, we note that the above procedures have been tested using Monte Carlo simulations, and
we find that the input mass profiles can be successfully recovered.
i ≃ (1 + x)2(δ log L/L∗)2 + (δ log Σcr,i)2 + δ2
Following optimization, we uniformly rescale the estimated errors by setting δscale so that
the best-fit model has χ2 = NDOF , the number of degrees of freedom. This does not alter the
optimized parameters, but does allow us to relate the uncertainties to the observed scatter in
the homology model. Moreover, it preserves the relative weighting among the data points, which
naturally gives more weight to quads and rings, and those systems with a measured zs. Rescaling is
particularly important because our model is undoubtedly a simplistic representation of galaxy mass
distributions. Hence, the χ2 is likely to be dominated by unmodeled complexity (i.e., deviations
from self-similarity) in the galaxy population, rather than by observational errors.
-- 12 --
5. Analysis and Results
We first simultaneously constrain all four parameters (Υ∗, x, fcdm, n) in our homology model.
The best-fit model has χ2 = 44.1 for NDOF = 18 (with δscale = 1). The rms scatter (0.15 in log M )
is significantly larger than can be accounted for by the assumed observational uncertainties. Given
our estimates of the measurement errors, the intrinsic scatter is roughly 30% in mass. Fig. 1 shows
the pairwise parameter constraints. The contours correspond to ∆χ2 = 1, 2.30, 4 and 6.17. Unless
otherwise noted, these and all subsequent ∆χ2 values involve the rescaled χ2, in which δscale is set
so that the best-fit model has χ2 = NDOF .
There are well-defined regions allowed by the data, but substantial degeneracy between param-
eters. This is expected because the mass-radius relations constrain the combined (luminous plus
dark matter) radial mass distribution. For example, we require a projected mass of ∼ 3 × 1011M⊙
inside 2Re for the typical L∗ lens galaxy. This can be achieved by placing more mass in the lumi-
nous component, which requires a larger Υ∗, or more mass in the dark matter component, which
requires a smaller Υ∗. This degeneracy is reflected in the Υ∗−fcdm panel of Fig. 1. The mass-radius
relations also prefer some overall shape, or concentration, for the combined profile. To achieve a
given mass concentration, we can place more of the mass in the stellar component if the CDM slope
is shallower, and less if the CDM slope is steeper. These degeneracies (Υ∗ − n, fcdm − n) are also
clearly seen in Fig. 1.
The slope x, which describes the increase in stellar mass-to-light ratio with luminosity, is
uncorrelated with the other three parameters (Fig. 1). Optimizing over the other parameters,
we find x = 0.14+0.16
−0.12 at 68% confidence (∆χ2 < 1), and −0.08 < x < 0.49 at 95% confidence
(∆χ2 < 4). This is the first constraint on x from strong lensing.
Our analysis robustly detects the presence of dark matter in the form of a mass component
that is more spatially extended than the light. Dark matter contributes several tens of percent
to the projected mass inside 2Re. This result can be quantified in a variety of ways, for different
assumptions:
• A model in which mass traces light (fcdm = 0) has ∆χ2 = 10.7 with respect to the overall
best fit, and is therefore rejected at > 99% confidence. It is too centrally concentrated to
account for the ensemble of aperture mass-radius relations. We note in passing that such a
model would require Υ∗ ≃ 11Υ⊙ in the rest frame B band, only slightly higher than the value
of Υ∗ = (7.8 ± 2.7)Υ⊙ determined by Gerhard et al. (2001) from dynamical analyses of the
central regions of local (nearly L∗) early-type galaxies.
• Optimizing over the other three parameters, we find fcdm > 0.36 at 68% confidence, and
fcdm > 0.08 at 95% confidence.
• For a shallow CDM density slope of n = 1, as suggested by the NFW profile, the best-fit
model has fcdm = 0.22, and ∆χ2 = 2.4 with respect to the best overall model. The allowed
-- 13 --
range is 0.12 < fcdm < 0.32 at 68% confidence, and 0.08 < fcdm < 0.45 at 95% confidence.
• For a steeper CDM density slope of n = 1.5, as suggested by the Moore et al. (1999b) profile,
the best-fit model has fcdm = 0.43, and ∆χ2 = 0.8 with respect to the best overall model.
The allowed range is 0.27 < fcdm < 0.57 at 68% confidence, and 0.14 < fcdm < 0.71 at 95%
confidence.
Because there are significant degeneracies between parameters, we cannot place a tight con-
straint on the slope of the dark matter component: 1.44 < n < 2.20 at 68% confidence, with an
upper bound of n < 2.33 at 95% confidence. However, the favored models have dark and luminous
components which sum to produce very similar mass profiles over the radial range probed by the
lensed images. This is illustrated in Fig. 2, which shows the projected mass M (R) versus R for all
models with ∆χ2 < 2.30 in the fcdm − n plane. We see that the mass profile closely tracks a pure
isothermal (M ∝ R) model.
In the scale-free limit (Υ∗ → 0 and fcdm → 1, see §4), we find n = 2.07±0.13 at 68% confidence,
consistent with isothermal or slightly steeper profiles. The 95% range is 1.83 < n < 2.33. The
best-fit scale-free model has ∆χ2 = 0.1 relative to the best overall model, so it provides a good
approximation to the mass distributions in lens galaxies.
Fig. 3 demonstrates how the ensemble of mass-radius relations traces the radial mass profile.
The solid line shows the scaled aperture mass M (R)/(L/L∗)1+x as a function of the scaled aperture
radius R/Re, predicted by the best-fit power-law model. Quad and ring systems are plotted at a
single radius REin, with enclosed mass ΣcrπR2
Ein. Doubles are more difficult to depict, as they
constrain a combination of two masses and radii. Note, however, that any mass profile can be
normalized such that the mass-radius relation for a specific double is satisfied exactly. The resulting
values of M (R1) and M (R2) in the context of the best-fit profile are plotted for each double in
Fig. 3. While the locations of the points obviously depend on the assumed model, the best-fit
profile will minimize their scatter about the model curve. In this way we see that doubles also help
map the mass profile.
The radial coverage offered by the current lens sample is not uniform (Fig. 3). Quad and ring
systems, which provide the strongest mass constraints, mostly lie in the range 1 . REin/Re . 4.
There are two notable outliers: Q2237+030 (Huchra et al. 1985), which probes the smallest radial
scale (REin/Re ≃ 0.2), and MG2016+112 (Lawrence et al. 1984), which probes the largest radial
scale (REin/Re ≃ 7). Because these lenses constrain the extremes of the density profile, it is
important to address how much they affect our fits. We find that dropping either of the systems does
not alter the primary conclusions that the total mass profile is nearly isothermal, or that x ≃ 0.14 is
favored. Some of the other constraints are slightly affected, and this can be understood based on the
unique characteristics of each lens. First, MG2016+112 probes a scale much larger than the effective
radius, and its aperture mass will be dominated by dark matter in most models. This system is
therefore important for constraining the properties of the CDM component. Removing it weakens
the exclusion of models in which mass traces light to about 95% confidence, and significantly
-- 14 --
weakens the upper limit on n in the limit fcdm → 0. Profile constraints in the scale-free limit
(fcdm → 1), however, are not significantly changed. Second, Q2237+030 probes a scale much
smaller than the effective radius, and its aperture mass will be dominated by luminous matter
in most models. Consequently, this system contributes minimally to constraints on the CDM
component. Dropping it has little effect on any of the model parameters, even those related to the
stellar mass.
Finally, it is interesting to dissect the distribution of scatter about the best-fit homology
model. Despite comprising only 7 of the 22 systems, doubles account for 56% of the χ2. The
additional scatter for doubles is greater than we would expect from the quadrupole smearing, and
may indicate some unmodeled effect which we do not presently understand. Dropping all doubles
would significantly improve the fit (χ2 = 17.1 for NDOF = 11, prior to rescaling) and tighten the
resulting parameter constraints, but would not alter the major conclusions. For example, a model
in which mass traces light would be rejected more strongly (∆χ2 = 19.6), and constraints on the
density slope in the scale-free limit would be n = 2.04 ± 0.09 (68% C.L.). We also note that a
pair of lenses are particularly strong contributors to the scatter. The two-image system Q0142 -- 100
(Surdej et al. 1987) is the farthest outlier, just as it is in our FP analysis (Rusin et al. 2003), and
accounts for almost 1/4 of the χ2. Because the galaxy appears to be much brighter than expected,
it greatly over-predicts aperture masses. The ring system MG1131+0456 (Hewitt et al. 1988) is
the second largest contributor to the χ2, accounting for 1/7 of its value, despite being assigned a
rather large fit tolerance due to its estimated source redshift. The lens galaxy can be brought into
better agreement with the other galaxies if the lensed source is at a substantially higher redshift
than the assumed median of zs = 2. While our sample is too small to reasonably remove these
two outliers, doing so would improve the fit (χ2 = 28.3 for NDOF = 16, prior to rescaling) but not
affect the major conclusions. For example, a model in which mass traces light would be rejected
somewhat more strongly (∆χ2 = 11.1), and constraints on the density slope in the scale-free limit
would be n = 2.08 ± 0.10 (68% C.L.).
6. Applications
6.1. Time Delays and the Hubble Constant
The time delays between gravitationally lensed images depend on a combination of the Hubble
constant and the mass distribution of the lensing galaxy. The Hubble constant inferred from a
measured time delay can be approximated as
H0 = A(1 − hκi) + Bhκi(η − 1) + C ,
(11)
where hκi ≡ hΣi/Σcr is the mean scaled surface mass density in the annulus defined by the radii
of the lensed images, η is the logarithmic slope of the density (κ ∝ R1−η) within the annulus,
and the constants A and B incorporate the lensing geometry, the redshifts, and the measured
-- 15 --
delay (Kochanek 2002). The A term is the most important, and the B term just contributes a
small (∼10%) correction. Kochanek (2002, 2003) investigates "simple" time delay lenses which are
dominated by a single galaxy with a precisely determined centroid relative to the images. Two
limiting cases of the mass distribution are considered: models in which mass traces light yield
h ≃ 0.7, while isothermal models yield h ≃ 0.5. A number of recent results from a broader
(perhaps less well understood) sample of lenses fall within the range 0.45 < h < 0.65 (Burud et al.
2002a, 2002b; Fassnacht et al. 2002; Treu & Koopmans 2002b; Winn et al. 2002) if isothermality
is assumed. In fact, the only lens to yield h ≃ 0.7 for an isothermal model is B0218+357 (Biggs
et al. 1999), but the result is currently dominated by systematics related to the poorly constrained
galaxy position. One can therefore make the case that estimates of the Hubble constant from
strong lensing are in broad agreement, but are systematically lower than the values favored by
both the HST Key Project (Freedman et al. 2001) and the recent WMAP results (Spergel et
al. 2003). Because Kochanek (2002, 2003) shows that nearly constant mass-to-light ratio models
are necessary to reconcile time delay measurements with h ≃ 0.7, this would seem to imply that
lens galaxies have no extended dark matter component, in contradiction to virtually every other
observational test.
We can now explore this intriguing problem in the context of our constraints on self-similar
galaxy models. Of the five lenses considered by Kochanek (2002), three are drawn from our sample.
The two exceptions are B1600+434 (Jackson et al. 1995), which has a late-type lens galaxy (Koop-
mans, de Bruyn & Jackson 1998) and is therefore not described by our models, and RXJ0911+0551
(Bade et al. 1997), which has a significant cluster perturbation complicating interpretation of the
image radii. Because PG1115+080 (Weymann et al. 1980) has been explored in detail by Treu
& Koopmans (2002b), we limit our demonstration to a pair of two-image lenses, SBS1520+530
(Chavushyan et al. 1997) and HE2149 -- 2745 (Wisotzki et al. 1996).
Because the favored homology models produce very similar mass profiles, hκi, η and the implied
H0 do not vary much along the fcdm − n degeneracy stripe. These three quantities are plotted
for SBS1520+530 and HE2149 -- 2745 in Figs. 4 and 5, respectively, by normalizing the models to
reproduce the mass-radius relations exactly. The values of H0 use the coefficients A, B and C from
Kochanek (2002). The uncertainties in these coefficients are propagated to H0 using Monte Carlo
procedures. Assuming that each model (fcdm,i,nj) has a likelihood pij ∝ exp(−∆χ2
ij/2), constraints
on H0 can be determined from a Bayesian analysis. These constraints are shown in panel (d) of
Figs. 4 and 5. For SBS1520+530, h = 0.58 ± 0.08 (68% C.L.). For HE2149 -- 2745, h = 0.55 ± 0.10
(68% C.L.). The values of h come out slightly higher than those obtained by Kochanek (2002) for
an isothermal profile, because our favored profiles are slightly steeper than n = 2. The Hubble
constants are still lower than those favored by the HST Key Project (h = 0.72 ± 0.08; Freedman et
al. 2001) and WMAP data (h = 0.72± 0.05; Spergel et al. 2003), but the discrepancy is only about
1σ in our analysis.
-- 16 --
6.2. Flux Ratio Anomalies and Microlensing
Simple, smooth models for the galaxy mass distribution have great difficulty reproducing the
observed flux ratios in four-image lenses (e.g., Dalal & Kochanek 2002; Metcalf & Zhao 2002; Chiba
2002; Keeton et al. 2003). Moreover, flux ratio measurements in both the radio and optical appear
to violate fundamental symmetry arguments which predict nearly equal magnifications for merging
image pairs. Such flux ratio "anomalies" are almost certainly due to small-scale structure in the
gravitational potential (Kochanek & Dalal 2003), which can significantly perturb magnifications
(e.g., Mao & Schneider 1998) while leaving the image positions virtually unaltered. There is a debate
regarding the identification of the perturbers. CDM simulations (Moore et al. 1999a) predict sig-
nificant substructure (with mass scale M & 106M⊙) in galaxy halos, and Dalal & Kochanek (2002)
demonstrate that the substructure mass fraction required to account for radio flux ratio anoma-
lies is in line with N -body results. Stars are another source of substructure in the gravitational
potential (Witt et al. 1995), and microlensing-induced time variability has been unambiguously
detected in the light curves of Q2237+030 (Wozniak et al. 2000). Schechter & Wambsganss (2002)
have pointed out that there is a tendency for saddle point images to be demagnified compared to
our expectations for smooth lens models (see also Kochanek & Dalal 2003). They demonstrate
that stellar microlensing produces this effect only if the stars account for a relatively small fraction
(0.15 . κlum/κ . 0.30) of the surface mass density at the image radius.8 If κlum/κ is too high,
then the asymmetric effects on saddles and local minima disappear; if κlum/κ is too low, then the
anomalies are too rare. We can now test whether our two-component mass models fall within the
favored range.
In the context of our homology model, the value of κlum/κ at a given radius R/Re depends only
on the dark matter abundance parameter fcdm and density slope n. Fig. 6a plots κlum/κ at 2Re,
a typical radial distance for lensed images. Since we require models with an extended dark matter
component, lower values of κlum/κ are favored at 2Re due to the fact that the surface brightness of
the galaxy has already decreased substantially at this radius, while the dark matter density has not.
Assuming that each model (fcdm,i,nj) has a likelihood pij ∝ exp(−∆χ2
ij/2), constraints on κlum/κ
can be determined from a Bayesian analysis. These constraints are shown in Fig. 6b. At Re, just
about any value of κlum/κ is permitted. At 2Re, however, the results are quite restrictive, favoring
κlum/κ < 0.31 (68% C.L.). Hence, our models are consistent with the values of κlum/κ needed to
produce the relative demagnification of saddle point images observed in flux ratio anomalies.
8To ward off possible confusion of terminology, we note that while fcdm is a global parameter describing the
projected dark matter mass fraction inside 2Re, κlum/κ is a local parameter describing the fraction of surface mass
density in the form of stars at the radius of a lensed image.
-- 17 --
7. Discussion
We have constrained the typical mass profile of early-type lens galaxies using an ensemble of
aperture mass-radius relations from 22 gravitational lenses. The different galaxies are combined
using a self-similar mass model consisting of four parameters: the present-day (z = 0) stellar
mass-to-light ratio in the B band for an L∗ galaxy (Υ∗), its dependence on galaxy luminosity (x,
where Υ ∝ Lx), the projected mass fraction of dark matter inside 2 effective radii (fcdm), and
the logarithmic density slope of the CDM (n, where ρ ∝ r−n). Despite significant parameter
degeneracies, the favored models have dark and luminous components which sum to produce very
similar total mass profiles over the radial range probed by the lensed images. In the scale-free limit
we find n = 2.07 ± 0.13 (68% C.L.), consistent with isothermal or slightly steeper profiles. Two-
component models imply that the dark matter mass fraction is not very high: 0.12 < fcdm < 0.32
for an n = 1 halo (NFW 1997), or 0.27 < fcdm < 0.57 for a steeper n = 1.5 halo (Moore et al.
1999b), both at 68% confidence. Even so, the need for an extended dark matter component in
the inner regions of the galaxies is significant, as a model in which mass traces light is ruled out
at > 99% confidence. These are the best statistical constraints yet obtained on the typical radial
mass profile in gravitational lens galaxies, and they add to evidence that the mass distribution is
nearly isothermal on the scale of a few effective radii.
One might expect lenses to be a biased sample of early-type galaxies, because deflectors with
more centrally-concentrated profiles have larger lensing cross sections per unit mass. Consequently,
if the galaxy population exhibits a range of profiles, then lensing would tend to select galaxies
with steeper mass distributions. Note, however, that profile constraints for the lens galaxy pop-
ulation are consistent with the nearly isothermal profiles favored by dynamical and X-ray studies
of local ellipticals, which are not selected on the basis of mass concentration. This would appear
to strengthen the argument that early-type galaxies do not exhibit much structural variety, and
therefore, that gravitational lenses are typical members of this galaxy population. Of course, even if
lenses were a biased galaxy sample, it would have no effect on applications such as Hubble constant
determination, which only requires an understanding of the mass profiles in lens galaxies.
Our results have implications for the relationship between dark and luminous matter in early-
type galaxies, and can be compared with recent studies involving the stellar dynamics of local
galaxies (Gerhard et al. 2001) and lenses at intermediate redshift (KT), the number and size
distribution of lenses (Keeton 2001), and the existence of a tight fundamental plane (Borriello,
Salucci & Danese 2003). While the use of different mass models and priors makes a detailed
comparison difficult, we can nonetheless survey the basic results for consistency.
First, consider evidence for the existence and abundance of dark matter within a few optical
radii. Our results rule out a galaxy population in which mass traces light at > 99% confidence,
and imply a small but significant dark matter mass fraction inside 2Re. KT likewise reject con-
stant mass-to-light ratio models based on the measured velocity dispersions of lens galaxies.
In
comparison, Gerhard et al. (2001) claim 10 -- 40% dark matter within the volume defined by Re, but
-- 18 --
caution that some elliptical galaxies show no evidence for dark matter at this scale. Coming from
the other side, Keeton (2001) uses a variety of lensing statistics to place an upper limit of 40% dark
matter within the volume defined by 2Re. Borriello et al. (2003) also suggest that a cuspy dark
matter halo can contribute little to the mass inside a few effective radii, if early-type galaxies are to
occupy a tight fundamental plane. We can compare these measurements to the predictions of the
CDM model. Assuming an NFW halo profile, a cooled baryon fraction of fcool ≡ Ωb,cool/ΩM ≃ 0.02
(from the local census of baryons by Fukugita, Hogan & Peebles 1998), and a typical initial halo
concentration (c ≃ 9; Bullock et al. 2001), the adiabatic compression models of Keeton (2001)
predict fcdm ≃ 0.6 -- 0.8. The predicted value of fcdm can be reduced slightly by increasing the
cooled baryon fraction, but even maximally efficient cooling (fcool = Ωb/ΩM = 0.17; Spergel et al.
2003) yields fcdm ≃ 0.4 -- 0.5. Because the above suite of measurements seem to imply lower dark
matter mass fractions in the central regions of early-type galaxies, they may support the evidence
from late-type galaxies (e.g., McGaugh & de Blok 1998; de Blok et al. 2001; Salucci 2001; de Blok
& Bosma 2002) that halos are less concentrated than predicted by numerical simulations of CDM
(e.g., Moore et al. 1999b; Bullock et al. 2001).
−0.12
Next consider the dependence of mass-to-light ratio on luminosity. We find that Υ ∝ L0.14+0.16
(68% C.L.), the first such constraint from strong lensing. While this result strictly applies to the
stellar mass-to-light ratio, the self-similarity of our model means that it also applies to the total
mass-to-light ratio inside any fixed fraction of the effective radius. The luminosity dependence
is slightly shallower than the slope of 0.5 . x . 0.7 measured by Gerhard et al. (2001) in the
B band using stellar dynamics. Each of these results is broadly consistent with the value of
x ≃ 0.30 ± 0.05 obtained by interpreting the B-band FP under the assumption of structural
homology (e.g., Jorgensen et al. 1996). Similar investigations have recently been performed at
longer wavelengths. Bernardi et al. (2003) measure x = 0.14 ± 0.02 in the r∗ band using virial
mass-to-light ratios, and derive a similar value from their FP slopes. In addition, Borriello et al.
(2003) show that scatter in the local r-band FP is minimized by x ≃ 0.2. Finally, it is interesting
to note that the above estimates of x, determined at the scale of the optical radius, are similar
to the relations between virial mass and luminosity determined from weak lensing and satellite
dynamics: Guzik & Seljak (2002) and Prada et al. (2003) find Mvir ∝ L1.2−1.6, while McKay et al.
(2001, 2002) favor Mvir ∝ L. The broad consistency of these scaling laws suggests that the mass
distribution in early-type galaxies is closely related to the light over many orders of magnitude in
radius.
We have applied our constraints on the mass distribution of early-type galaxies to a pair of
interesting problems. First, we find that the measured time delays in SBS1520+530 and HE2149 --
2745 favor a Hubble constant of h ≃ 0.55 − 0.60. This value is slightly higher than that derived by
Kochanek (2002) using isothermal models, since we prefer mass profiles that are slightly steeper
than n = 2. The derived H0 is still systematically lower than recent results (Freedman et al. 2001;
Spergel et al. 2003), although the discrepancy is only about 1σ in our analysis, as the statistical
uncertainties in the mass profile are still significant. Second, we have estimated the fraction of
-- 19 --
the surface mass density in the form of stars (κlum/κ) at the radii of lensed images. For images
at 2Re, our models favor κlum/κ < 0.31 (68% C.L.). This is consistent with the values at which
stellar microlensing can reproduce the observed flux ratio anomalies, in particular the tendency for
saddle-point images to be more strongly perturbed than minima (Schechter & Wambsganss 2002).
Much more work is needed to fully understand the structure of early-type galaxies, and the
implications for galaxy formation theories. Because of its unique advantages, gravitational lensing
is certain to contribute significantly to this goal. Improved statistical tests to constrain the mass
profiles of early-type galaxies require larger samples of lenses. Particularly important is the addition
of new quad and ring systems which probe the mass distribution on small (REin < Re) and
large (REin > 4Re) scales, and therefore fill in the tails of our radial coverage. As always, the
key to turning recently discovered systems into useful astrophysical tools is the acquisition of
high-quality HST photometry and ground-based spectroscopy. Expanded samples could allow us
to replace our simplistic power-law model for the dark matter halo with more realistic profiles,
and consider structural dependencies on luminosity or color.
In this way we may be able to
properly investigate sources of scatter in our homology model, and perhaps quantify or constrain
the structural diversity of the early-type galaxy population. With regard to lensing determinations
of the Hubble constant, it is vital to increase the number of lenses that have both direct profile
constraints and well-determined time delays. Currently there is only one such system (PG1115+080;
Schechter et al. 1997; Treu & Koopmans 2002b).9 Programs are now underway to both measure
more time delays, and obtain profile constraints on current systems by measuring stellar velocity
dispersions or mapping extended emission from the lensed host galaxies. Such observations should
greatly illuminate current discrepancies related to time delays and the Hubble constant, and, more
generally, improve our understanding of the mass distribution in early-type galaxies.
We thank Josh Winn and the anonymous referee for offering comments and suggestions which
greatly improved this manuscript. We acknowledge the support of HST grants GO-7495, 7887, 8175,
8804, and 9133. We acknowledge the support of the Smithsonian Institution. CSK is supported by
NASA ATP Grant NAG5-9265.
9A number of profile constraints have been claimed for the time delay lens Q0957+561 (e.g., Grogin & Narayan
1996), but Keeton et al. (2000) demonstrates that existing models fail to properly reproduce lensed extended emission
from the quasar host galaxy.
-- 20 --
REFERENCES
Bade, N., Siebert, J., Lopez, S., Voges, W., & Reimers, D. 1997, A&A, 317L, 13
Bernardi, M., Renzini, A., da Costa, L.N., Wegner, G., Alonso, M.V., Pellegrini, P.S., Rit´e, C., &
Willmer, C.N.A. 1998, ApJL, 508, L143
Bernardi, M., et al. 2003, AJ, 125, 1866
Bertin, G., et al. 1994, A&A, 292, 381
Bertin, G., Ciotti, L., & Del Principe, M. 2002, A&A, 386, 149
Biggs, A.D., Browne, I.W.A., Helbig, P., Koopmans, L.V.E., Wilkinson, P.N., & Perley, R.A. 1999,
MNRAS, 304, 349
Blumenthal, G.R., Faber, S.M., Flores, R., & Primack, J.R. 1986, ApJ, 301, 27
Borriello, A., Salucci, P., & Danese, L. 2003, MNRAS, 341, 1109
Bradac, M., Schneider, P., Steinmetz, M., Lombardi, M., King, L.J., & Porcas, R. 2002, A&A, 338,
373
Bruzual, A.G., & Charlot, S. 1993, ApJ, 405, 538
Bullock, J.S., Kolatt, T.S., Sigad, Y., Somerville, R.S., Kravtsov, A.V., Klypin, A.A., Primack,
J.R., & Dekel, A. 2001, MNRAS, 321, 559
Burud, I., et al. 2002a, A&A, 383, 71
Burud, I., et al. 2002b, A&A, 391, 481
Caon, N., Capaccioli, M., & D'Onofrio, M. 1993, MNRAS, 265, 1013
Chavushyan, V.H., Vlasyuk, V.V., Stepanian, J.A., & Erastova, L.K. 1997, A&A, 318L, 67
Chiba, M., 2002, ApJ, 565, 17
Ciotti, L., Lanzoni, B., & Renzini, A. 1996, MNRAS, 282, 1
Cohn, J.D., Kochanek, C.S., McLeod, B.A., & Keeton, C.R. 2001, ApJ, 554, 1216
Dalal, N., & Kochanek, C.S. 2002, ApJ, 572, 25
de Blok, W.J.G, McGaugh, S.S., Bosma, A., & Rubin, V.C., 2001, ApJL, 552, L23
de Blok, W.J.G., & Bosma, A. 2002, A&A, 385, 816
Djorgovski, S., & Davis, M. 1987, ApJ, 313, 59
Dressler, A., Lynden-Bell, D., Burstein, D., Davies, R.L., Faber, S.M., Terlevich, R., & Wegner, G.
1987, ApJ, 313, 42
Evans, N.W., & Hunter, C. 2002, ApJ, 575, 68
Fabbiano, G. 1989, ARA&A, 27, 87
Faber, S.M., Dressler, A., Davies, R.L., Burstein, D., & Lynden-Bell, D. 1987, in: Nearly normal
galaxies, ed. S.M. Faber (New York: Springer), 175
-- 21 --
Fassnacht, C.D., Xanthopoulos, E., Koopmans, L.V.E., & Rusin, D. 2002, ApJ, 581, 823
Ferreras, I., Charlot, S., & Silk, J. 1999, ApJ, 521, 81
Finch, T.K., Carlivati, L.P., Winn, J.N., & Schechter, P.L. 2002, ApJ, 577, 51
Fischer, P., et al. 2000, AJ, 120, 1198
Freedman, W.L., et al. 2001, ApJ, 553, 47
Fukugita, M., Hogan, C.J., & Peebles, P.J.E. 1998, ApJ, 503, 518
Gerhard, O., Kronawitter, A., Saglia, R.P., & Bender, R. 2001, AJ, 121, 1936
Grogin, N.A., & Narayan, R. 1996, ApJ, 464, 92
Guzik, J., & Seljak, U. 2002, MNRAS, 335, 311
Hewitt, J.N., Turner, E.L., Schneider, D.P., Burke, B.F., & Langston, G.I. 1988, Nature, 333, 537
Holder, G.P., & Schechter, P.L. 2003, ApJ, 589, 688
Holtzman, J.A., Burrows, C.J., Casertano, S., Hester, J.J., Trauger, J.T., Watson, A.M., &
Worthey, G. 1995, PASP, 107, 1065
Huchra, J., Gorenstein, M., Kent, S., Shapiro, I., Smith, G., Horine, E., & Perley, R. 1985, AJ, 90,
691
Im, M., et al. 2002, ApJ, 571, 136
Jackson, N., et al. 1995, MNRAS, 274L, 25
Jorgensen, I., Franx, M., & Kjaergaard, P. 1996, MNRAS, 280, 167
Keeton, C.R., Kochanek, C.S., & Seljak, U. 1997, ApJ, 482, 604
Keeton, C.R., et al. 2000, ApJ, 542, 74
Keeton, C.R., Christlein, D., & Zabludoff, A.I. 2000, ApJ, 545, 129
Keeton, C.R. 2001, ApJ, 561, 46
Keeton, C.R., Gaudi, B.S., & Petters, A.O. 2002, ApJ, submitted (astro-ph/0210318)
Keeton, C.R. 2003, ApJ, 582, 17
Kleinheinrich, M., et al. 2003, astro-ph/0304208
Kochanek, C.S. 1995, ApJ, 445, 559
Kochanek, C.S., et al. 2000, ApJ, 543, 131
Kochanek, C.S., Keeton, C.R., & McLeod, B.A. 2001, ApJ, 547, 50
Kochanek, C.S., & White, M. 2001, ApJ, 559, 531
Kochanek, C.S. 2002, ApJ, 578, 25
Kochanek, C.S. 2003, ApJ, 583, 49
Kochanek, C.S., & Dalal, N. 2003, ApJ, submitted (astro-ph/0302036)
-- 22 --
Koopmans, L.V.E., de Bruyn, A.G., & Jackson, N. 1998, MNRAS, 295, 534
Koopmans, L.V.E., & Treu, T. 2002, ApJL, 568, L5
Koopmans, L.V.E., & Treu, T. 2003, ApJ, 583, 606
Lawrence, C.R., Schneider, D.P., Schmidet, M., Bennett, C.L., Hewitt, J.N., Burke, B.F., Turner,
E.L., & Gunn, J.E. 1984, Sci., 223, 46
Leh´ar, J., Langston, G.I., Silber, A., Lawrence, C.R., & Burke, B.F. 1993, AJ, 105, 847
Leh´ar, J., et al. 2000, ApJ, 536, 584
Loewenstein, M., & White, R.E. 1999, ApJ, 518, 50
Madgwick, D.S., et al. 2002, MNRAS, 333, 133
Mao, S., & Schneider, P. 1998, MNRAS, 295, 587
Matsushita, K., Makishima, K., Ikebe, Y., Rokutanda, E., Yamasaki, N., & Ohashi, T. 1998, ApJL,
499, L13
McGaugh, S.S., & de Blok W.J.G. 1998, ApJ, 499, 41
McKay, T.A., et al. 2001, ApJ, submitted (astro-ph/0108013)
McKay, T.A., et al. 2002, ApJL, 571, L85
Metcalf, R.B., & Madau, P. 2001, ApJ, 563, 9
Metcalf, R.B., & Zhao, H.S. 2002, ApJL, 567, L5
Moore, B., Ghigna, S., Governato, F., Lake, G., Quinn, T., Stadel, J., & Tozzi, P. 1999a, ApJL,
524, L19
Moore, B., Quinn, T., Governato, F., Stadel, J., & Lake, G. 1999b, MNRAS, 310, 1147
Munoz, J.A., Kochanek, C.S., & Keeton, C.R. 2001, ApJ, 558, 657
Navarro, J.F., Frenk, C.S., & White, S.D.M. 1997, ApJ, 490, 493
Pahre, M.A., de Carvalho, R.R., & Djorgovski, S.G. 1998, AJ, 116, 1606
Prada, F., et al. 2003, ApJ, submitted (astro-ph/0301360)
Renzini, A., & Ciotti, L. 1993, ApJL, 416, L49
Rix, H.-W., de Zeeuw, P.T., Cretton, N., van der Marel, R.P., & Carollo, C.M. 1997, ApJ, 488, 702
Romanowsky, A.J., & Kochanek, C.S. 2001, ApJ, 553, 722
Rubin, V.C., Thonnard, N., & Ford, W.K., Jr. 1980, ApJ, 238, 471
Rusin, D., & Ma, C.-P. 2001, ApJL, 549, L33
Rusin, D., & Tegmark, M. 2001, ApJ, 553, 709
Rusin, D., et al. 2001, AJ, 122, 591
Rusin, D., Norbury, M., Biggs, A.D., Marlow, D.R., Jackson, N.J., Browne, I.W.A., Wilkinson,
P.N., & Myers, S.T. 2002, MNRAS, 330, 205
-- 23 --
Rusin, D., et al. 2003, ApJ, 587, 143
Saha, P., & Williams, L.L.R. 2001, AJ, 122, 585
Salpeter, E. 1955, ApJ, 121, 161
Salucci, P. 2001, MNRAS, 320L, 1
Schade, D., et al. 1999, ApJ, 525, 31
Schechter, P.L., et al. 1997, ApJL, 475, L85
Schechter, P.L., & Wambsganss, J. 2002, ApJ, 580, 685
Schneider, P., Ehlers, J., & Falco, E.E. 1992, Gravitational Lenses (Berlin: Springer-Verlag)
Spergel, D.N., et al. 2003, ApJ, submitted (astro-ph/0302209)
Surdej, J., Swings, J.-P., Magain, P., Courvoisier, T.J.-L., & Borgeest, U. 1987, Nature, 329, 695
Tonry, J.L. 1998, AJ, 115, 1
Treu, T., Stiavelli, M., Bertin, G., Casertano, S., & Moller, P. 2001, MNRAS, 326, 237
Treu. T., Stiavelli, M., Casertano, S., Moller, P., & Bertin, G. 2002, ApJL, 564, L13
Treu, T., & Koopmans, L.V.E. 2002a, ApJ, 575, 87
Treu, T., & Koopmans, L.V.E. 2002b, MNRAS, 337L, 6
Treu, T., & Koopmans, L.V.E. 2003, MNRAS, in press (astro-ph/0306045)
van Albada, T.S., Bertin, G., & Stiavelli, M. 1995, MNRAS, 276, 125
van de Ven, G., van Dokkum, P.G., & Franx, M. 2002, MNRAS, submitted (astro-ph/0211566)
van Dokkum, P.G., Franx, M., Kelson, D.D., & Illingworth, G.D. 1998, ApJL, 504, L17
van Dokkum, P.G., Franx, M., Kelson, D.D., & Illingworth, G.D. 2001, ApJL, 553, L39
Weymann, R.J., et al. 1980, Nature, 285, 641
Williams, L.L.R., & Saha, P. 2000, AJ, 119, 439
Winn, J.N., Kochanek, C.S., McLeod, B.A., Falco, E.E., Impey, C.D., & Rix, H.-W. 2002, ApJ,
575, 103
Winn, J.N., Rusin, D., & Kochanek, C.S. 2003, ApJ, 587, 80
Wisotzki, L., Koehler, T., Lopez, S., & Reimers, D. 1996, A&A, 315L, 405
Witt, H.J., Mao, S., & Schechter, P.L. 1995, ApJ, 443, 18
Wozniak, P.R., Alard, C., Udalski, A., Szymanski, M., Kubiak, M., Pietrzynski, G., & Zebrun, K.
2000, ApJ, 529, 88
Zaritsky, D., Smith, R., Frenk, C.S, & White, S.D.M. 1997, ApJ, 478, 39
This preprint was prepared with the AAS LATEX macros v5.0.
-- 24 --
Table 1. Structural Parameters for Lens Systems
Lens
zd
zs Morph.
Radii (′′)
REin = 1.35
REin = 1.19
REin = 0.71
3.60 Quad
2.72 Double R1 = 0.38, R2 = 1.86
2.64 Quad
1.34 Quad
3.12 Double R1 = 0.61, R2 = 2.22
1.54 Double R1 = 0.18, R2 = 1.39
2.32 Double R1 = 1.10, R2 = 2.09
1.72 Quad
Ring
2.81 Quad
3.40 Quad
3.62 Quad
1.86 Double R1 = 0.39, R2 = 1.21
1.17
Ring
1.39 Quad
Ring
1.74
Ring
3.27 Quad
Quad
REin = 1.15
REin = 1.14
REin = 1.05
REin = 0.50
REin = 1.63
REin = 1.14
REin = 1.15
REin = 1.05
REin = 0.86
REin = 1.42
REin = 0.78
0047 -- 2808
Q0142 -- 100
MG0414+0534
B0712+472
HS0818+1227
B1030+074
HE1104 -- 1805
PG1115+080
MG1131+0456
HST14113+5211
HST14176+5226
B1422+231
SBS1520+530
MG1549+3047
B1608+656
MG1654+1346
B1938+666
MG2016+112
B2045+265
HE2149 -- 2745
Q2237+030
B2319+051
0.49
0.49
0.96
0.41
0.39
0.60
0.73
0.31
0.84
0.46
0.81
0.34
0.72
0.11
0.63
0.25
0.88
1.00
0.87
0.50
0.04
0.62
--
--
--
2.03 Double R1 = 0.34, R2 = 1.35
1.69 Quad
REin = 0.88
--
Double R1 = 0.61, R2 = 0.82
Note. -- Listed for each lens are the galaxy (zd) and source (zs)
redshifts, morphology, and relevant radii (Einstein radius REin for
quads and rings; image radii R1 and R2 for doubles). Uncertainties
on these radii are negligible. The image radii for B2319+051 are
the estimated values from Rusin et al. (2001).
-- 25 --
Fig. 1. -- Constraints on the self-similar model. We show pairwise constraints on the present-day
(z = 0) stellar mass-to-light ratio in the B band for an L∗ galaxy (Υ∗, in solar units), its logarithmic
dependence on luminosity (x), the projected CDM mass fraction inside 2 effective radii (fcdm), and
the logarithmic CDM density slope (n). Solid contours represent ∆χ2 = 2.30 and 6.17, the 68%
and 95% confidence levels for two parameters. Dotted lines represent ∆χ2 = 1 and 4, the 68% and
95% confidence levels for one parameter. The errors have been rescaled so that the best-fit model
has χ2 = NDOF .
-- 26 --
Fig. 2. -- Mass profiles for allowed two-component homology models. Each model falls within the
68% (∆χ2 < 2.30) confidence region of the fcdm − n plane (Fig. 1). Solid lines are the projected
masses inside R/Re, where Re is the effective radius. Profiles are normalized to a fixed projected
mass at R = 2Re. For comparison we show the de Vaucouleurs profile (dotted line), and an offset
isothermal profile (dashed line). While the allowed models exhibit a wide range of dark matter
abundances, they all have total mass profiles which are approximately isothermal over the radial
range spanned by the lensed images.
-- 27 --
Fig. 3. -- The aperture mass profile in the power-law limit. We plot the scaled projected mass
M (R)/(L/L∗)1+x as a function of the scaled radius R/Re. The solid line is the prediction of
the best-fit power-law model with n = 2.07 and x = 0.14. Data points represent individual lens
systems. Solid squares are lenses with measured source redshifts; open squares are lenses with
estimated (zs = 2.0 ± 1.0) source redshifts. For quads and rings, a single point is plotted showing
REin and mass ΣcrπR2
Ein. For doubles (big circles), points are plotted at each image radius (R1
and R2) and connected with a dotted line, and the masses are the M (R1) and M (R2) which would
satisfy the mass-radius relation exactly in the context of this profile (see §5).
-- 28 --
Fig. 4. -- Interpreting the time delays of SBS1520+530. The first three panels show various
quantities as a function of CDM mass fraction fcdm and CDM density slope n. Solid contours
indicate (a) the mean scaled surface density hκi inside the image annulus, (b) the effective density
slope (η) inside the image annulus, and (c) the derived Hubble constant H0. The models are
normalized to satisfy the mass-radius relation exactly. Note that the quantities exhibit the expected
behavior: η → n for fcdm → 1 and hκi = 0.5 for a scale-free isothermal model. Also, we see that an
isothermal profile yields H0 ≃ 50 km s−1 Mpc−1, as demonstrated by Kochanek (2002). The dotted
contours show the 68% (∆χ2 < 2.30) and 95% (∆χ2 < 6.17) confidence regions from the homology
model (Fig. 1). Plotted in panel (d) is the relative probability of H0 based on the information
displayed in (c). A Hubble constant of H0 = (58 ± 8) km s−1 Mpc−1 is favored (68% C.L.).
-- 29 --
Fig. 5. -- Interpreting the time delays of HE2149 -- 2745. The plots are analogous to those in Fig. 4.
A Hubble constant of H0 = (55 ± 10) km s−1 Mpc−1 is favored (68% C.L.).
-- 30 --
Fig. 6. -- The fraction of the surface mass density in the form of stars. (a) Plotted are contours
of κlum/κ at R = 2Re for models in the fcdm − n plane. The dotted contours show the 68%
(∆χ2 < 2.30) and 95% (∆χ2 < 6.17) confidence regions from the homology model (Fig. 1). (b)
The relative probability of κlum/κ based on the information displayed in (a). The solid line is the
fraction at R = 2Re. The dashed line is the fraction at R = Re.
|
astro-ph/9803115 | 1 | 9803 | 1998-03-10T18:44:44 | The Evolution of Blue Stragglers Formed Via Stellar Collisions | [
"astro-ph"
] | We have used the results of recent smoothed particle hydrodynamic simulations of colliding stars to create models appropriate for input into a stellar evolution code. In evolving these models, we find that little or no surface convection occurs, precluding angular momentum loss via a magnetically-driven stellar wind as a viable mechanism for slowing rapidly rotating blue stragglers which have been formed by collisions. Angular momentum transfer to either a circumstellar disk (possibly collisional ejecta) or a nearby companion are plausible mechanisms for explaining the observed low rotation velocities of blue stragglers. Under the assumption that the blue stragglers seen in NGC 6397 and 47 Tuc have been created solely by collisions, we find that the majority of these blue stragglers cannot have been highly mixed by convection or meridional circulation currents at anytime during their evolution. Also, on the basis of the agreement between the predictions of our non-rotating models and the observed blue straggler distribution, the evolution of blue stragglers is apparently not dominated by the effects of rotation. | astro-ph | astro-ph | The Evolution of Blue Stragglers Formed Via Stellar Collisions.
J. A. Ouellette and C. J. Pritchet
Department of Physics & Astronomy,
University of Victoria,
Box 3055, Victoria BC, V8W 3P6 Canada
Electronic mail: [email protected], [email protected]
ABSTRACT
We have used the results of recent smoothed particle hydrodynamic simu-
lations of colliding stars to create models appropriate for input into a stellar
evolution code. In evolving these models, we find that little or no surface convec-
tion occurs, precluding angular momentum loss via a magnetically-driven stellar
wind as a viable mechanism for slowing rapidly rotating blue stragglers which
have been formed by collisions. Angular momentum transfer to either a cir-
cumstellar disk (possibly collisional ejecta) or a nearby companion are plausible
mechanisms for explaining the observed low rotation velocities of blue stragglers
Under the assumption that the blue stragglers seen in NGC 6397 and 47 Tuc
have been created solely by collisions, we find that the majority of blue stragglers
cannot have been highly mixed by convection or meridional circulation currents
at anytime during their evolution. Also, on the basis of the agreement between
the predictions of our non-rotating models and the observed blue straggler dis-
tribution, the evolution of blue stragglers is apparently not dominated by the
effects of rotation.
8
9
9
1
r
a
M
0
1
1
v
5
1
1
3
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
-- 2 --
1.
Introduction
Blue stragglers lie roughly along an extension of a star cluster's main sequence
(MS) and are generally bluer and brighter than the turn-off (TO) stars. While
several possible mechanisms for their formation have been proposed (e.g. delayed
star formation, binary mass transfer, binary coalescence, stellar collisions; see
Livio, 1993, and Stryker, 1993, for reviews), the collision scenario, in which ini-
tially unbound stars come into contact and merge during a dynamical encounter,
has received the most attention recently (e.g. Sandquist et al. 1997; Ouellette
& Pritchet 1996; Sills, Bailyn & Demarque 1995). Blue stragglers formed by
collisions are of particular interest because their formation rate tells us about the
stellar interaction rate in their cluster environment and the dynamical history of
the cluster itself (Bailyn, 1995).
There are at least two ways to investigate the formation rate of blue strag-
glers by collisions. One is to perform N-body simulations, modelling the stellar
environment and determining the collision rate directly. This requires knowledge
of the stellar interaction cross-section, local stellar density, mass function, and
binary fraction. The second is to model the structure of the merger remnants
and to evolve them using a stellar evolution code. Just as the age of a cluster
is found by comparing theoretical isochrones, based on standard stellar models,
with an observed colour-magnitude diagram (CMD), comparison of the evolution
of the merger models to the observed numbers and distribution of blue stragglers
in the CMD can tell us about the distribution of lifetimes, and the production
rates, of these stars.
In this paper, we develop and evolve models of collisional merger remnants
and compare the predictions of these models with the observed blue straggler
distributions in NGC 6397 and 47 Tuc.
2. Predictions of the Collision Scenario -- A Review.
There is direct and indirect evidence that blue stragglers are more massive
than the turn-off (TO) stars in their parent clusters (spectroscopic measurements:
Shara et al. 1997, Rodgers & Roberts 1995; mass segregation: e.g. Sarajedini
& Forrester 1995, Lauzeral et al. 1992, Mathieu & Latham 1986), suggesting a
formation mechanism which involves combining the mass of at least two stars into
one massive star (e.g. Livio 1993). Although there are a few possible mechanisms
by which this can be achieved, this paper is concerned with only one of them
-- the collision scenario. This requires that two stars come into direct contact
during a strong dynamical encounter and then merge, leaving behind a massive
-- 3 --
remnant. As well as being more massive, the collisionally merged star will have
other properties which we will summarize here.
Hills and Day (1976) investigated the possibility that collisions could have oc-
curred between single stars in globular clusters (GCs) within the lifetimes of the
clusters. They found that, in the dense cores of some GCs, the timescale for sin-
gle star interactions is short enough that a significant number of collisions would
have occurred. Also, because the impact parameter at the time of the collision
is random, the collisions will, on average, occur off-axis. Angular momentum
conservation then requires that the merged star be rapidly rotating. Hills & Day
hypothesised that the excess energy from the collision would mix nuclear pro-
cessed material throughout the merger remnant, as well as mixing hydrogen into
the core, essentially "resetting the nuclear clock" to the chemically homogeneous
zero-age main sequence (ZAMS) state.
Hoffer (1983) and Leonard (1989) extended the analysis of collision probabil-
ities to include binary-single star and binary-binary interactions. Because of the
increased physical cross-section (essentially the semi-major axis of the binary)
and the additional gravitational focussing due to the larger mass of the bound
pair, the probability that a collision will occur is greatly increased if one or more
of the interacting objects is a binary. Since the rate at which binary interactions
will occur in a population of stars will increase with the binary fraction, we ex-
pect to see a significant collision rate in clusters with a large binary fraction and
in the cores of clusters, where most of the binaries are expected to be found due
to mass segregation.
Leonard and Fahlman (1991) found that the rate of production of blue strag-
glers in NGC 5053 needed to be at least 1 blue straggler every ∼2.5×108years,
assuming that the average lifetime of a blue straggler is ∼6×109years, to main-
tain the observed numbers. From their scattering experiments, a strong binary-
binary interaction should occur every ∼2.2×107years, with only 1 in 20-40 such
interactions actually resulting in a stellar collision (hence, one blue straggler pro-
duced every ∼(4 − 8)×108 years). According to Sigurdsson & Phinney (1993),
the timescale for exchanges during strong interactions is significantly shorter than
the timescale for collisions. Also, such exchanges tend to produce a net hardening
of the binary, which will lengthen the timescale for collisions.
Leonard (1996) noted that the high frequency of binaries observed among
blue stragglers (e.g. Kaluzny et al. 1997, Kaluzny 1997, Edmonds et al. 1996)
could be accounted for by binary-binary interactions. A dynamical encounter
that is strong enough to result in a direct collision between two of the component
stars is also likely to result in a third star being captured into an eccentric orbit
-- 4 --
around the newly formed blue straggler; the fourth star is ejected from the system,
removing the excess binding energy.
These dynamical studies treated the interacting stars as point sources, and
assumed that a direct collision would take place and form a blue straggler if two of
the binary components approached within some factor of their radii, ignoring the
dynamics of the stellar collision itself. Benz & Hills (1987, 1992) studied stellar
collisions by performing smoothed particle hydrodynamic (SPH) simulations of
colliding n = 3/2 polytropes. They found that the resulting merger remnant's
properties depended on the relative mass of the parent stars. During collisions
involving equal mass stars, Benz & Hills found that the material from the parent
stars was distributed throughout the merger remnant, resulting in a star with
a roughly homogeneous chemical composition: essentially a ZAMS star. Their
collisions involving unequal mass stars resulted in a merger remnant with the
material from the less evolved, low mass parent star in the core of the remnant;
the material from the high mass star was, in contrast, mixed throughout.
In
either case, the merger remnant roughly resembled a ZAMS star due to the
increased hydrogen content in the core.
Lombardi, Rasio & Shapiro (1996, hereafter LRS) found that the high degree
of mixing which Benz & Hills had observed in their merger remnants was largely
an artifact of the low resolution of their simulations and their choice of n = 3/2
polytropes as approximations to GC stars. LRS noted that evolved GC stars
would be more centrally condensed than n = 3/2 polytropes and so chose to
use n = 3 polytropes to approximate the structure of TO stars, and n = 3/2
polytropes for lower mass stars. SPH simulations using these approximations
demonstrated that there should be little or no hydrodynamical mixing during a
stellar collision.
Benz & Hills (1987, 1992) and LRS found that significant mass loss could
occur during grazing collisions; in the high impact velocity collisions of Benz &
Hills, the two stars could be completely disrupted, leaving no single remnant
behind.
If the star was not disrupted during the collision, it was swollen to
pre-main sequence sizes by the injection of orbital kinetic energy into the stellar
material. The greatest amount of mass loss for a given impact velocity occurred
during off-axis collisions.
The high rotation velocity which the collision scenario predicts for blue strag-
glers is one of its potential weak points.
In open clusters, such as M67 (Pe-
terson et al. 1984, Mathys 1991, Pritchet & Glaspey 1991), blue stragglers are
not observed to be rapidly rotating. The only reported observation of the ro-
tation velocity of a blue straggler in a globular cluster (Shara et al. 1997),
-- 5 --
V sin i = 155 ± 55km/s, is high, but not necessarily unusual for normal stars
of the same spectral type (A7V: V rot∼130 − 170km/s, Lang 1992) nor is it ro-
tating near to its break-up velocity of ∼ 410km/s (estimated using the mass and
log g from Shara et al.). Leonard & Livio (1995) suggested that the rotation
velocity of a blue straggler could be lowered by angular momentum loss (AML)
via a magnetically driven stellar wind, similar to the mechanism proposed for
pre-main sequence stars by Tout & Pringle (1992). For this mechanism to work,
the blue straggler would need to be left in a largely convective, Hayashi-like phase
after the collision.
In addition to providing an AML mechanism for blue stragglers, a highly
convective phase would also mix the stars, homogenizing them despite the fact
that they will be initially inhomogeneous (LRS). Bailyn & Pinsonneault (1995)
and Sills, Bailyn & Demarque (1995) found that such a high degree of mixing
is needed to explain the colours and luminosities of blue stragglers seen in some
globular clusters. On the other hand, Ouellette & Pritchet (1996) found that blue
stragglers tend to avoid the ZAMS and that a high degree of mixing during a
pre-main sequence phase is essentially ruled out for most blue stragglers by their
distribution on the CMD. If significant surface convection occurs, though not
necessarily enough to mix the star, abundance anomalies may be observable. In
fact, many of the blue stragglers in M67 have been observed to be underabundant
in lithium (Pritchet & Glaspey, 1991) and have anomalous CNO abundances
(Mathys, 1991), possibly indicative of mixing.
In short, the properties of blue stragglers formed through collisions are that
they should be more massive than the cluster TO stars, they might be rapidly
rotating, they might have surface chemical anomalies, and they are likely to have
a companion in an eccentric orbit.
3. Structure of Merger Remnants.
Scattering experiments, such as those of Leonard (1989) and Sigurdsson &
Phinney (1993), tell us at what rate we can expect stellar collisions to occur.
However, in order to test the accuracy and viability of the collision scenario,
we need to know the timescales over which collisional merger remnants will be
observable as blue stragglers. This can be done by creating stellar models from
the predictions of the SPH simulations of stellar collisions, and following their
evolution with a stellar evolution code. Once these timescales are known, they
can be combined with the results of scattering experiments to predict the number
of blue stragglers and other remnants of such strong interactions (Bailyn, 1995)
-- 6 --
which should be observable in a cluster. Also, calculation of the evolution of
merger remnants allows us to make predictions concerning the distribution of
blue stragglers in the CMD. Comparison of this distribution with observations
allows us to infer whether collisional mergers are likely to have occurred in a
population of blue stragglers, and even the dynamical history of the cluster itself.
As mentioned earlier, LRS have studied stellar collisions by performing SPH
simulations of colliding polytropes. Their results showed that little, if any, hy-
drodynamical mixing occurred during the collision. The composition profiles of
their merger remnants can be understood as an 'entropy stratification' of the
stellar material during the collision: the gas from the parent stars settles in the
merger remnant such that the final entropy profile increases from the core out-
wards. This entropy stratification can result in some unusual chemical profiles
(as shown by LRS), but also allows some prediction of the chemical profile of a
collisional remnant -- which we will use in this study.
The collisions studied by LRS were between equal mass polytropes and un-
equal mass polytropes, for a variety of different masses. We choose here to model
only mergers between equal mass stars and mergers between a TO star and a
lower mass star. Throughout the rest of this paper, we refer to these merger
events as "equal mass mergers" and "unequal mass mergers", respectively. Al-
though there are any number of possible combinations of parent star masses which
would result in a particular merger mass between 1MT O and 2MT O, the mergers
considered here represent the extremes of hydrogen content: equal mass mergers
will result in the highest possible hydrogen content for a particular merger mass,
whereas unequal mass mergers will result in the lowest possible hydrogen content.
3.1. Predictions from Entropy Stratification.
In their simulations, LRS approximated MS TO stars by n = 3 polytropes
while lower mass stars were approximated by n = 3/2 polytropes or compos-
ite polytropes. Their initial polytropic models had entropy and density profiles
similar to those shown in Figure 1. During a collision involving a n = 3 poly-
trope (∼0.80M⊙, MS TO star) and a n = 3/2 polytrope (M ∼<0.40M⊙, lower MS
star), entropy stratification predicts that the material of the lower mass, and
presumably less evolved, n = 3/2 polytrope will settle into the core of the merger
remnant, bringing with it a fresh supply of hydrogen. The subsequent evolu-
tion of the merger remnant will be strongly affected by the amount of hydrogen
brought into the core by the entropy stratification. This stratification of the ma-
terial also provides a simple explanation of why no nuclear processed material is
-- 7 --
brought to the surface of the merger remnant during a collision.
If the entropy of the stellar gas is not modified during a collision, the dis-
tribution of the parent stars' material throughout the merger remnant can be
found using the entropy profiles in Figure 1 -- this leads directly to the merger
remnant's chemical profile if those of the parent stars are also known. However,
shock heating during the collision can modify the entropy of the gas depending
on the dynamics of the collision and the form of viscous dissipation chosen for
the SPH (LRS). During relatively gentle collisions, such as the head-on, parabolic
collisions studied by LRS, little shock heating will occur and so the entropy of
the gas at the time of the collision can be used to determine the final merger
profile (See section 3.1.2).
3.1.1. Stellar Collisions versus Polytropic Collisions.
During the previous discussion we have stressed that the simulations of LRS
and Benz & Hills (1987, 1992) describe polytropic collisions. The reason for this
pointed emphasis is that stars, especially evolved ones in GCs, are not polytropes.
The structure of a polytrope is given by
P (r) = Kρ(r)(1+1/n)
(1)
where P is the pressure, ρ is the gas density, n is the polytropic index and K
is a constant. Using this relationship as a constraint, ρ(r) and K can be found
using only two of the equations of stellar structure: the equations of hydrostatic
equilibrium and mass continuity. For a simple polytrope, the molecular weight
µ is constant throughout, in which case both K and n are also constants. This
allows the structure of the polytrope to be determined with no information about
sources of opacity or nuclear processes.
The entropy profile of a polytrope can be found by assuming an ideal gas
equation of state of the form
P = AρΓ1,
where Γ1 is one of the adiabatic exponents and is defined by
dP
P
+ Γ1
dV
V
= 0.
(2)
(3)
If radiation pressure within the gas can be ignored, then Γ1 = γ = cP /cV =
5/3 for an ideal, monatomic gas. Equating the pressure from the polytropic
structure to the pressure from the equation of state yields A throughout the
gas. The relationship between the entropy S and the constant A can be found
-- 8 --
by integrating the first law of thermodynamics, T dS = dU + P dV , using the
definition of γ and Equation 3. This yields
S =
N0k
µγ
lnA + const
(4)
where N0 is Avagadro's number and k is Boltzmann's constant. From the form
of the polytropic structure equation (Eq. 1) and the equation of state (Eq. 2), if
1 + 1/n is not equal to γ, then A, and hence the entropy S, will not be constant
throughout the star. Polytropes with n = 3/2 will have 1 + 1/n = γ and so will
have constant entropy.
Polytropes are reasonable approximations to ZAMS stars since they, like poly-
tropes, are chemically homogeneous, or at least nearly so. As pointed out by LRS,
n = 3 polytropes approximate the structure of radiative stars while n = 3/2
polytropes have a structure similar to that of fully convective stars. However,
although GC TO stars are radiative, they are far from being chemically homo-
geneous. Figure 2 shows the entropy and density profiles for a GC TO star and
lower MS stars for a cluster like 47 Tuc. The differences between the entropy and
density profiles of GC stars and the polytropes shown in Figure 1 are largely due
to the chemical inhomogeneity of the GC stars. The increased molecular weight
toward the centre of an evolved star results in a lower pressure at a fixed density
and temperature, requiring the star to adjust its internal structure to maintain
hydrostatic equilibrium. Equation 4 shows that this increase in the molecular
weight results in a decrease in the entropy of the gas.
The decrease of the entropy in the core of an evolved star will directly affect
the final chemical profile of a merger found through entropy stratification. From
Figure 2, it is obvious that no material from a low mass star will be able to
penetrate to the core of a MS TO star, leaving the merger remnant with a helium-
rich core, unlike the collision between two equivalent polytropes. The evolution
of two such merger remnants would be completely different if no mixing takes
place in a subsequent convective phase of evolution.
Comparison of Figures 1 and 2 shows that an n = 3 polytrope underestimates
the central density of a TO star by more than a factor of ten. This in itself is
enough to argue that no hydrogen-rich material from a lower mass star will
penetrate to the core of a merger involving a TO star and a lower MS star (an
unequal mass merger in our terminology). Entropy stratification of the material
during such a collision will produce a merger remnant with a dense, helium-rich
core.
A collision involving equal mass stars or one involving equal mass polytropes
-- 9 --
should produce a merger remnant with a composition profile nearly identical to
those of the parent stars.
3.1.2. Shock Heating.
As mentioned earlier, shock heating can modify the entropy of the gas during
the collision and so affect the distribution of material throughout the merger
remnant.
If the amount of shock heating is significant, the structure of the
merger remnant could differ considerably from what would be expected if the
unmodified entropy profiles of the parent stars were used to estimate the final
structure.
The amount of shock heating due to the passage of a shock can be estimated
from the pre-shock state of the gas, and the velocity of the shock. Making
the simple assumptions that the sound velocity within the gas is equal to the
adiabatic sound speed, c2
∂ρ , and that there is no change of state due to the
passage of the shock (i.e. γf inal = γinitial) then, taking the Mach number of the
shock as M0(= VS
cs ), the change in entropy after the passage of the shock is (e.g.
Shore 1992)
s = ∂P
ln" P1
P0! ρ1
ρ0!−γ#
∆S = S1 − S0 =
N0k
µγ
(5)
where
and
ρ1
ρ0
=
(γ + 1)M 2
0
(γ − 1)M 2
0 + 2
P1
P0
=
2γM 2
− (γ − 1)
0
γ + 1
Here, the subscript 0 refers to the conditions on the immediate pre-shock side
of the shock discontinuity, and the subscript 1 refers to those conditions imme-
diately on the post-shock side of the discontinuity. Figure 3 shows the effect
of shock heating on the entropy profiles of the stars shown in Figure 2 under
the assumption that the shock velocity remains constant throughout the star.1 It
is obvious that there is potentially significant shock heating near the surface of
the stars and very little in the cores. In particular, entropy stratification of the
parent stars' material using the 'post-shock' entropy profiles in Figure 3 would
1This also neglects the effects of viscous dissipation, which would affect both ∆S and the velocity of the
shock. The initial shock velocity is taken to be the impact velocity (See Sec. 4). Typically, the shock has a
Mach number ∼ 5 at M (r) ∼ 0.99M∗, and ∼< 2 at the centre.
-- 10 --
produce a very similar stratification to that found using the pre-shock profiles.
The change in entropy near the surface of the star may result in some mixing
of the surface layers, however, since these layers should contain no nuclear pro-
cessed material, this would not significantly affect the subsequent evolution of
the merger remnant.
The amount of shock heating calculated above is undoubtably an underes-
timate -- an actual collision would result in multiple shocks, especially during
grazing collisions, and would have a much more complex geometry than that
used in the above calculation. A more accurate estimate of the amount of shock
heating could be found through simulations using SPH, which has been shown
to model shocks well, depending on the form of artificial viscosity used (Mon-
aghan 1992). Despite the rough nature of this estimate of the amount of shock
heating, the relative amounts of shock heating should be correct: that is, the en-
velopes of the stars are more affected by the shocks than the interiors, and that
the amount of shock heating is not enough to greatly affect the stratification of
nuclear processed material when dealing with dissimilar stars.
3.2. Physical Structure of Merger Remnants.
Leonard & Livio (1995) have suggested that a blue straggler formed through
a stellar collision would resemble a pre-main sequence star immediately after
the collision. The SPH simulations of LRS show that the remnant of a collision
involving two polytropes can be several times larger than either of the parent
stars; however, unlike a pre-main sequence star, the central density of the remnant
will be much higher than that found on the Hayashi track. The decrease in central
density of polytropic mergers relative to the parent stars is generally less than
50%, and is as little as ∼10% in head-on mergers (from the simulations of LRS).
Because of the size and dense core of a merger remnant, it will more closely
resemble a red giant branch star than a star on the Hayashi track.
4. Construction of Initial Models.
Despite the facts that the SPH simulations of LRS use polytropes rather than
evolved stars, and that the differences between the two species of objects can be
extreme, we can use their results as the basis for producing models for input into
a stellar evolution code. Entropy stratification of the material, which is a direct
consequence of the dynamics of the fluid interaction and a requirement for the
SPH fluid stability (see LRS for a discussion), can be used to predict the chemical
-- 11 --
profile of the merger remnant. Although there is no equivalent procedure for
predicting the density profile of the remnant, the simulations of LRS show that
the central density will not, in general, be much lower than that of the parent
stars -- this can be used as an additional constraint when constructing models.
Since head-on, parabolic collisions are relatively gentle, they are also the
easiest to approximate for our purposes. We choose to ignore mass loss during
the collision, the effects of rotation and the departure from spherical symmetry it
produces. Because head-on collisions produce the least amount of mass loss and
lowest rotation rates, these approximations are reasonably valid. We also choose
to ignore shock heating during the collision and use the pre-collision entropy
profiles of our parent stars to define the chemical profile of the merger product
using entropy stratification. The amount of shock heating during a head-on
collision is small (Sec 3.1.2 and LRS) and so should not strongly affect the end
result.
The initial models we used to investigate blue straggler evolution were formed
in three steps from a series of standard stellar models of the appropriate age
and metallicity. First, for collisions between a TO star and a lower MS star,
an appropriate TO model was scaled in mass so that its mass agreed with the
total mass of the two parent stars; for equal mass mergers, the mass of one of
the parent stars was scaled. Next, the scaled model's composition profile was
replaced with the chemical profile found by entropy stratification of the parent
stars. This model was then expanded to simulate the expansion of the star during
the collision.
The expansion of the model was performed by adding an additional term ǫx
to the energy balance equation of stellar structure:
∂L
∂M
= ǫnuc − ǫν + ǫgrav + ǫx,
where ǫnuc is the nuclear energy generation rate, ǫν is the energy loss due to
neutrinos, and ǫgrav is the energy generation due to contraction of the star (=
−T ∂S
∂t ). If ǫx were held constant throughout the star, the central density would
decrease rapidly as the star expanded and the star would end up on, or near,
the Hayashi track (this is similar to the procedure which Sandquist et al. (1997)
used to produce their Hayashi models). As mentioned earlier, the density of the
core does not decrease by a large factor during the collision; to ensure this, we
assumed that ǫx falls off as 1/ρ (see below). During the expansion of our models,
the central density does not decrease by more than 20% which, as explained
below, results in our final models being virtually independent of the form of ǫx
after they have relaxed to the MS.
-- 12 --
One additional consideration is how much energy to inject into the star before
allowing it to contract to the MS and evolve normally. We have chosen to use a
criterion which takes into account the binding energy of the parent stars and the
kinetic energy of the collision. The binding energy of a star can be expressed as
Ebind = −q
GM 2
R
,
where M and R are the mass and radius of the star, and q is related to the degree
of central concentration (Cox & Giuli, 1968):
q = Z 1
0
M(r/R)/M
r/R
·
dM(r/R)
M
.
In the centre of mass frame of the two stars the orbital energy is
Eorb =
1
2
µv2 −
GMµ
r
,
which, for a parabolic orbit, is conveniently equal to zero.
(Here, M is the
total mass of the two stars, M1 + M2, and µ is the reduced mass of the system,
µ = M1M2
M1+M2
.) Hence, the kinetic energy of the orbit is
K =
GMµ
r
.
Assuming that one half of the orbital kinetic energy goes into expanding the
merger remnant, the final binding energy of the merger remnant is E ′
bind =
Ebind1 + Ebind2 + 1
2K. For the purposes of determining the kinetic energy at
the time of the collision, we have set the separation of the two stars r equal
to the average of their radii, r = R1+R2
. The expansion of the model is halted
when its binding energy equals E ′
bind. For the same reasons that the form of
ǫx is not critical, as long as the density constraint is obeyed (discussed below),
the structure of the blue straggler after it has contracted to its MS is not highly
dependent on the exact value of E ′
bind, although the duration of any convective
zones during the pre-main sequence evolution is affected.
2
One might be concerned about the ramifications for the models of assuming an
energy injection term proportional to 1/ρ. We have investigated this by adopting
different energy injection schemes and comparing the resultant models. We find
that, although the tracks that the models follow on the CMD as they expand and
contract can differ considerably, the final models that satisfy the above criteria for
energy and central density are very similar and evolve identically after they have
contracted to the MS. This can be explained using the Vogt-Russell Theorem,
-- 13 --
which states that the structure of a star in hydrostatic and thermal equilibrium
is uniquely determined by the total mass and the run of chemical composition
throughout the star (Vogt 1926, Russell 1927; see e.g. Cox & Giuli, 1968). Thus,
if the amount of mixing which takes place is not significantly affected by the form
of the energy injection term, the evolution of the merger remnant, at least after
it has contracted to the MS, should also be unaffected by the choice. Our tests
using different forms of ǫx show that this is the case as long as the central density
does not decrease beyond what is indicated by the SPH simulations.
As pointed out by the referee, our energy injection scheme is similar to that
used by Podsiadlowski (1996) in his investigation of the response of stars to
heating by tidal effects. In his exploratory paper, Podsiadlowski investigated the
effect of various forms of energy injection (e.g. centrally concentrated, uniform,
surface) on a 0.8M⊙ GC ZAMS star -- the different forms of energy injection
were intended to approximate the zones in which tidally excited oscillations might
dissipate their energy. Podsiadlowski found that the response of the star and
its structure after a fixed amount of energy had been injected were strongly
dependent upon the way in which energy was injected. Our initial experiments
into the various forms of ǫx were quite similar to Podsiadlowski's and lend support
to his findings; the differences in the final structures from our early experiments
and those of Podsiadlowski's were simply due to the fact that our initial models
were evolved, whereas his were chemically homogeneous. Unlike Podsiadlowski's
investigation, however, the form of ǫx for our models is constrained by the results
of SPH: the central density of head-on mergers does not decrease dramatically
from that of the parent stars'. From our experience, which is similar to what is
reported by Podsiadlowski, forms of ǫx which are uniform throughout the star
or are centrally concentrated result in a rapid decrease in the central density
-- by the time enough energy is injected to meet our energy constraint (E ′
bind
-- discussed earlier), the star's structure would resemble that of a star on the
Hayashi track.
4.1. Cluster Parameters.
We have chosen to compare our models of merger remnants with the blue
stragglers observed in NGC 6397 and 47 Tuc. Both clusters have high central
densities (see Table 1), a fact that makes the blue stragglers in these clusters
excellent candidates for formation by collisions.
The stellar evolutionary code used to produce and evolve our merger models
is a modified version of the VandenBerg et al. (1997) stellar evolution code and
-- 14 --
V
opacities. For each cluster, a grid of stellar models was produced, and isochrones
were extracted using the method of Bergbusch & VandenBerg (1992). Ages and
TO masses were found for each cluster by matching the TO luminosity with
these isochrones; absolute TO luminosities were found by using the "best-fit"
M HB
-[Fe/H] relationship of Chaboyer et al. (1996) to obtain a distance modulus
for each cluster. The ages derived in this manner agree with those derived by
Chaboyer et al., and are shown in Table 1 along with the other derived cluster
parameters. It should be noted that, although the ages and TO masses derived
here are dependent upon the M HB
-[Fe/H] relationship or distance calibration
chosen, the final characteristics of our models (other than mass) are only weakly
dependent upon the choice -- in particular, the surface convection seen in the
models presented here becomes less effective if the distance scale is increased.
V
5. Results and Discussion.
Blue straggler models were produced as described in the previous sections for
both unequal mass and equal mass mergers, and for masses up to twice the cluster
TO mass. Figures 4 and 5 show the evolution of these models on the CMD. As
would be expected, because of the chemical inhomogeneity of the merger remnant
due to entropy stratification, the stars tend to avoid the ZAMS; only the low
mass, relatively unevolved, equal mass mergers approach what would appear to
be normal ZAMS stars. The differences between the tracks for the mergers of
NGC 6397 and 47 Tuc are largely due to the difference in cluster metallicity.
These differences, including convection and a comparison with the observations,
will be discussed in the next sections.
5.1. Surface Convection.
Leonard & Livio (1995) and Sills, Bailyn & Demarque (1995) have suggested
that blue stragglers must become largely convective during their post-collision,
pre-main sequence phase of evolution to explain both the observed rotation ve-
locities and colours. While the blue straggler models of Sandquist et al. (1997)
lend some support to this, our models do not, nor do the similar models of Sills
et al. (1997).
Sandquist et al. (1997) performed SPH simulations of mergers of equal mass
polytropes and evolved the products of these mergers by imposing the resultant
composition profile on to standard stellar models of the same mass as the mergers,
and then forcing the stars to expand until they had reached the Hayashi track.
-- 15 --
Their models developed deep surface convection, enough to bring at least some
helium to the surface of the stars. However, forcing the stars onto the Hayashi
track would decrease the central density beyond what is observed to occur in SPH
simulations (LRS) -- thus requiring an additional energy source which would not
be present in the actual collision. In addition, the deep convective envelopes seen
by Sandquist et al. are a consequence of the fact that their models are forced on
to the Hayashi track, at which point the low surface temperatures will require
surface convection. Surface convection will persist until the opacity in the outer
regions of the star decreases to the point where radiative transfer becomes efficient
-- either when the star has evolved to higher surface temperatures, or convection
has brought a significant amount of helium to the surface.
Sills et al. (1997) created initial blue straggler models directly from the en-
tropy and density profiles produced by SPH. None of their models developed
any surface convection until they had evolved onto the red giant branch. The
total lack of surface convection during the pre-main sequence phase prevents any
helium-rich material from being dredged up to the surface layers of the star,
and makes spin-down of a rapidly rotating blue straggler by a magnetic wind
mechanism (Leonard & Livio, 1995) implausible. However, because the outer
portions of SPH merger remnants are not necessarily in dynamical equilibrium,
Sills et al. found it necessary to extrapolate the outer structure of their merger
models from the purely radiative interior, forcing the envelope to be radiative,
at least initially.
Surface convection does occur in some of our models during the pre-main
sequence phase, unlike the models of Sills et al., but not to the extent found
by Sandquist et al.. As in standard models evolving to the MS, the depth and
duration of the surface convection zones in our models depends on the star's mass.
The most massive merger remnants (M ∼1.8 − 2.0 × MT O), whether formed by
an unequal mass or equal mass merger, never develop surface convection zones
during the phases prior to reaching the MS. The convection zones seen in our
lower mass models are generally deeper and longer lived in the lowest mass stars,
decreasing in depth and duration for mergers of higher mass. The convective
zones typically contain ∼<4% of the star's mass and usually last for less than a
few times 106years.
5.1.1. Consequences of Surface Convection and Angular Momentum Loss.
The thin, short-lived surface convection seen in our models, and the absence
of surface convection in the models of Sills et al. (1997), precludes any wind-
-- 16 --
driven AML mechanism for slowing down the rapidly rotating blue stragglers
predicted by the collision scenario. However, blue stragglers are not observed to
be rapidly rotating in open clusters where, despite the comparatively low stellar
density, it is possible for a fraction of blue stragglers in open clusters to be formed
by collisions (Leonard 1996, Leonard & Linnell 1992). To account for the low
rotation velocities, there must be either an additional AML mechanism acting
which is not dependent upon surface convection, or the blue stragglers in open
clusters are being formed by a mechanism other than collisions which might
produce more slowly rotating blue stragglers (e.g. binary coalescence, binary
mass transfer), in which case the estimated numbers of collisionally generated
blue stragglers in open clusters is incorrect. Even if the production of blue
stragglers by collisions is ruled out in open clusters other formation mechanisms
can be assisted, or accelerated, by strong dynamical interactions (Sigurdsson &
Phinney, 1993).
Little is known about the rotation of blue stragglers in GCs, although the
stage has been set to rectify this problem by the observations of Shara et al.
(1997). BSS-19 (Paresce et al., 1991) in 47 Tuc is estimated to have a mass
of 1.70 ± 0.40M⊙,2 compared to the cluster TO mass of ∼0.86M⊙, and a high
rotation velocity (V sin i = 155 ±55km/s). From our models, this star should not
have had a convective envelope at any time during its "pre-blue straggler" phase,
assuming it was created by a collisional merger, and so should contain the same
angular momentum with which it was created, unless an AML mechanism other
than a magnetically driven wind has acted upon it. During its contraction to the
MS, its rotational velocity should have increased from its initial value by a factor
of ∼5 − 6, due to angular momentum conservation, implying an initial rotation
velocity of ∼ 30km/s. From the results of LRS, this would have required a nearly
head-on collision which, although not unlikely, is less probable than an off-axis
collision. Hence, it is more probable that BSS-19, if created by a collision, is either
inclined close to the line of sight (sin i ∼< π
4 ), has experienced some AML, or was
instead formed through some mechanism other than a collision, as suggested by
Shara et al..
It is possible that AML from an initially rapidly rotating blue straggler could
be achieved by angular momentum transfer to a circumstellar disk, possibly ejecta
from the collision, or to a nearby companion, possibly captured during a binary
2Comparing the observations of effective temperature (Tef f = 7630 ± 300K) and surface gravity (log g =
4.09 ± 0.1) directly to our models yields a mass estimate of 1.55 ± 0.10M⊙, independent of distance, in
reasonable agreement with the determination of Shara et al..
-- 17 --
interaction. Cameron et al. (1995) found that angular momentum transfer to a
circumstellar disk is an extremely efficient mechanism for slowing the rotation
of stars as they contract to the main sequence. This mechanism does require
that the star have a convective envelope for the generation of a magnetic field,
but the field strength required to shed a given amount of angular momentum is
not necessarily as high as that needed for a wind-driven mechanism. Angular
momentum transfer to a nearby companion has the advantage that a convective
envelope is not required and demands that many blue stragglers be in binary
systems, as is observed. During the distended contraction phase of the blue
straggler's evolution, a nearby companion could exert a considerable torque on
the star, forcing the stars to approach tidal lock. The angular momentum transfer
to the companion will tend to force it into a larger orbit and reduce any initial
eccentricity in the orbit, as well as slowing the rotation of the blue straggler.
5.1.2. Surface Abundances.
Although it is questionable whether surface convection is sufficient to explain
the moderate rotation rates of most blue stragglers, any amount of surface con-
vection could act to alter the surface abundances of these stars. Since we find
that surface convection does not occur in the most massive blue stragglers, it is
possible that abundance anomalies might be a way in which to distinguish be-
tween the formation mechanisms, as suggested by Sills et al. (1997) -- but only
for the most massive stragglers. The convective zones in some of our lower mass
models (M ∼< 1.40 × MT O) can penetrate to depths where the temperature is
high enough to destroy the more fragile elements, such as lithium (Pritchet &
Glaspey 1991, Hobbs & Mathieu 1991, Glaspey et al. 1994). However, although
the amount of hydrodynamical mixing which occurs during a collision appears to
be small, sufficient mixing should occur in the envelope during the merger (due
to shock heating, for example -- Sec 3.1.2) that some chemical anomalies might be
expected, even in the absence of convection. Additionally, meridional currents,
which should occur to some extent in rapidly rotating stars, will also mix the
surface layers even if the core is not penetrated (Tassoul & Tassoul, 1984).
5.2. Core Convection.
In the normal main sequence stars of NGC 6397 and 47 Tuc, core convection
should persist for the entire main sequence lifetime of stars with masses greater
than ∼1.40M⊙ and ∼1.20M⊙, respectively. Core convection does not occur at
-- 18 --
anytime in our unequal mass mergers, whereas core convection appears in most
of our equal mass models computed for 47 Tuc and for a short period of time
in one equal mass merger model computed for NGC 6397. No core convection
occurs during the distended pre-main sequence phase, but rather starts when the
model has contracted close to its main sequence.
That no core convection occurs in our unequal mass mergers is not surprising
due to the high central density and helium abundance. This is in contrast to
the results of Sills et al. (1997) who find that unequal mass mergers typically do
develop core convection. The simulations of Sills et al., however, are based on
models produced using the endproducts of polytropic collisions, which, as shown
earlier in the discussion on entropy stratification, will tend to have hydrogen-rich
cores. The increased hydrogen abundance in the core results in a higher central
opacity, making the central layers convectively unstable.
The differences in the amount of convection seen between the models for NGC
6397 and 47 Tuc are a consequence of the fact that NGC 6397 has a metallicity
which is ∼15 times lower than that of 47 Tuc, reducing the efficiency of the CNO
cycle for masses less than ∼1.40M⊙. At that mass our equal mass mergers for
NGC 6397 (MT O ∼ 0.71M⊙) have sufficiently low central hydrogen content that
convective transport is not necessary.
5.3. Consequences of Calculated Mixing Scales.
According to the Vogt-Russell Theorem, the structure of a star in hydrostatic
and thermal equilibrium is determined by its mass and composition profile, which
maps into a point on the H-R diagram. We would expect that a star perturbed
slightly from its position of equilibrium on the H-R diagram would relax back
to the same equilibrium position and structure. If the star is perturbed from its
equilibrium state to a greater extent, such that no additional convection occurs
which might change its chemical profile significantly, and such that no mass is
lost during the perturbation, we would still expect it to relax back to the same
equilibrium state on a timescale equal to the star's Kelvin-Helmholtz time scale.
Similarly, two merger remnants produced in collisions involving identical sets of
stars and having identical masses and chemical profiles, but which are initially
at different points in the H-R diagram, should produce identical blue stragglers
if no significant mixing occurs during their pre-main sequence phase.
This same argument was used earlier to explain why the exact details of
the mechanism used to expand our blue straggler models to their initial pre-
main sequence position are unimportant. However, here it has the additional
-- 19 --
implication that two theoretical merger remnant models which have identical
masses and chemical profiles, but were produced using different assumptions
about their structure, should produce identical blue stragglers if no significant
convective mixing occurs during their pre-main sequence phase. For this reason,
we expect the evolution of the models of Sills & Lombardi (1997) and of Sills et al.
(1997) to be similar to the evolution of our own unequal mass merger models and
equal mass merger models, respectively.
5.4. Comparison With Observations.
Shown on each of the tracks in Figures 4 and 5 are equally spaced intervals
of 0.05Gyr. Since the probability of seeing a star in any phase of its evolution
is proportional to the amount of time it spends in that phase, a blue straggler
which was created at some random time in the past has an equal probability of
being seen in any of the marked intervals on the appropriate evolutionary track.
This implies that the observed blue straggler distribution should cluster roughly
where the marked intervals are closest if the blue stragglers have been created
by processes similar to those represented by the models.
Figures 6 and 7 show the observed blue straggler distributions in NGC 6397
(Lauzeral et al., 1992) and 47 Tuc (Guhathakurta et al., 1992) compared with
the expected distribution 3 of blue stragglers from our equal and unequal mass
mergers.
Ignoring, for the moment, those blue stragglers which lie blueward
of the ZAMS, the blue straggler distributions in both clusters lie roughly in the
region predicted by our unequal mass mergers; the blue stragglers formed through
equal mass mergers tend to cluster too near to the ZAMS. For our unequal mass
mergers, the observed and expected distributions are similar at a 68% and 96%
confidence level for NGC 6397 and 47 Tuc, respectively; the same comparison
for our equal mass mergers yields 18% and 2% confidence levels (see Ouellette
& Pritchet 1998b for details). The obvious interpretation of this is that most of
3 The expected distribution was found by extracting stars from the tracks at random, using probability
distributions created from the evolution timescales of the tracks themselves. The two distributions of blue
stragglers, the observed and expected distributions, have the same distribution of apparent masses, excluding
the extreme blue stragglers as noted in the text. By this we mean that a comparison of the photometric
positions of the blue stragglers and our tracks yields, for each blue straggler, a probability distribution for
the mass. The fake blue stragglers have masses drawn from this probability distribution. The distribution
of fake blue stragglers also takes into account the expected photometric errors as indicated on the figures.
The probabilities quoted in the text are from Monte Carlo simulations which will be described in Ouellette
& Pritchet (1998).
-- 20 --
the blue stragglers in these two clusters may have been created through unequal
mass mergers or a similar event which leaves the remnant in an apparently evolved
state after formation (binary coalescence?).
There is, however, an additional, more critical observation to be made from
the apparent agreement between the distribution of the blue stragglers and the
predictions of our unequal mass merger models. The collisional scenario for the
formation of blue stragglers predicts that such mergers will, on average, be born
rapidly rotating. This rapid rotation can affect the evolution of the star by
providing non-thermal pressure support, which can extend the MS lifetime of a
star, or by initiating meridional circulation currents, which could mix the star.
Non-thermal pressure support would extend the lifetime of a star by requiring
less energy to be liberated from the core in the form of nuclear reactions, thereby
prolonging the hydrogen-burning lifetime of the star. Additionally, the star will
appear cooler and less luminous than its non-rotating equivalent (Clement, 1994).
Meridional circulation currents are a consequence of the distortion of the star's
gravitational potential by the rotation; as a result of this distortion, circulation
currents will be initiated and mix the star. If hydrogen-rich material is brought
into the core of the star by the circulation currents, the hydrogen-burning lifetime
of the star, and so its lifetime as an observable blue straggler, will be greatly
extended. Also, due to the increased central hydrogen abundance, the position
of the blue straggler on the CMD should be closer to the ZAMS than it would
be otherwise.
If blue stragglers created by collisions are rotating rapidly enough for these
effects to be important, the agreement between the predicted blue straggler dis-
tribution from our tracks, which do not include any effects of rotation, and the
observed blue straggler distribution should be poor. The apparent agreement on
the CMD between the prediction of the unequal mass mergers and the observed
blue stragglers places limits on the importance of the effects of rotation. First,
although the hydrogen burning lifetimes of the blue stragglers might in fact be
different than what our models predict, the difference cannot be extreme (cer-
tainly much less than a factor of two) else the observed blue straggler distribution
would be significantly more clustered near the MS. Second, if meridional circu-
lation were to mix the star, the observed blue stragglers would have significant
scatter blueward toward the ZAMS. This does not exclude meridional circula-
tion in the envelope (e.g. Tassoul & Tassoul, 1984), but little or no helium-rich
material will be brought to the surface in such a case.
-- 21 --
5.4.1. Blue Straggler Formation Rates.
A simple estimate of the formation rate of blue stragglers in these clusters
can be made by finding the average amount of time our merger models will spend
as observable blue stragglers. Assuming, as suggested by the comparison of the
observed and expected blue straggler distributions, that the blue stragglers in
these clusters are formed by unequal mass mergers, the average formation rates
of blue stragglers in NGC 6397 and 47 Tuc are 1 blue straggler every ∼5.0×107yr
and ∼3.6×107yr, respectively. For comparison, if the blue stragglers in these
clusters were formed solely by equal mass mergers the rates would be 1 blue
straggler every ∼2.3×108yr and ∼1.2×108yr. Assuming that the binary fraction
in the core of each cluster is 100% (following Leonard & Fahlman, 1991), and
that the average binary semi-major axis is ∼ 0.1AU, equation 4.12 of Sigurdsson
& Phinney (1993) yields an average time between collisions of ∼ 6×107yr and
∼ 7×105yr for NGC 6397 and 47 Tuc, respectively. These collision timescales
are close to lower limits, since the binary fraction is most likely not 100%; for
the timescale for 47 Tuc to agree with the model predictions, the binary fraction
would have to be ∼10%, or the average binary separation would have to be
smaller.
(A more detailed analysis will be presented in Ouellette & Pritchet
1998b).
5.5. Extreme Blue Stragglers.
In the above discussion and analysis we have pointedly ignored the blue strag-
glers which lie blueward of the ZAMS. While it is possible that some of these 'ex-
tremely blue' blue stragglers are merely field objects4 or photometric errors, many
other studies of clusters with excellent photometric quality and low foreground
object contamination (e.g. Bolte 1992) exhibit such extreme blue stragglers as
well. This suggests that these blue stragglers are produced by a different process
than that modelled here. In fact, as other studies have shown (Sandquist et al.
1997, Sills et al. 1997, Ouellette & Pritchet 1996), these stars lie roughly where
we would expect fully mixed merger remnants to be. However, inspection of
the 'fake' blue straggler distributions in Figures 6 and 7 shows that photometric
scatter may be responsible for some of these stars.
Earlier we suggested that meridional currents produced by rapid rotation are
4Guhathakurta et al. (1992) estimate that ∼< 3 of the blue stragglers they observed in 47 Tuc can be
explained by background objects in the SMC.
-- 22 --
not necessarily important when modelling blue stragglers produced by collisions.
It may in fact be that some rotational threshold exists above which this process
will become important for the more rapidly rotating blue stragglers.
Mestel (1953) pointed out that, although the average rotation rate of early
type (A-F) stars is rapid enough that one would expect meridional circulation
to be ongoing within these stars, the fact that we see such stars evolved to the
giant phase implies that they have chemically inhomogeneous structures, which
would not be the case if meridional circulation currents were present. Mestel
(and Tassoul & Tassoul, 1984) found that meridional currents were strongly in-
hibited by even a slight chemical gradient throughout the star. Since the stars in
a GC are evolved and so not chemically homogeneous, and we expect chemical
gradients in merger remnants, this may be enough to disrupt circulation cur-
rents in the majority of blue stragglers. Tassoul & Tassoul (1984) found that,
for stars rotating well below rotational break-up, the velocity of the circulation
currents increased with increasing rotational velocity, but the depth of penetra-
tion of the currents was virtually unchanged -- hence it is very unlikely that
meridional currents could be invoked to mix fresh hydrogen to the cores of these
stars. The predicted circulation in the envelope may still be enough to affect ob-
served chemical abundances, although it is questionable whether or not enough
helium will be mixed to the surface to cause the star to appear as an extreme
blue straggler. However, it may be possible that, if a star was rotating near to
rotational break-up, the meridional currents might be able to penetrate further
into the star than in Tassoul & Tassoul's more slowly rotating models -- this
would have to be confirmed through additional theoretical studies into rotation
using two-dimensional stellar models and is purely speculative at this point. If
meridional currents are responsible for mixing helium to the surface of extreme
blue stragglers, then these stars are most likely those which were created with
the highest initial rotation velocities.
6. Conclusions.
We have evolved stellar models which are appropriate for stars created during
collisions between equal and unequal mass stars. In doing so we have used the
results of SPH simulations of colliding polytropes, most importantly the predic-
tions of entropy stratification. To produce these models, we have made several
assumptions to facilitate the incorporation of the results of the SPH simulations
of collisions between polytropic stars into models of real stars; these assumptions
include neglecting mass loss, rotation, and shock heating during the collision. Be-
-- 23 --
cause we have restricted our attention to head-on parabolic collisions, we believe
these approximations to be reasonably valid; mass loss and rotational velocity
are shown to be small in this case by the simulations of LRS; shock heating
is shown to be small and, most importantly, should not affect the distribution
of helium-rich gas in the merger remnant (Sec 3.1.2), although it is likely that
shock heating will affect the distribution of material near the surface of the star.
The form of the energy injection term ǫx which we use to expand our merger
remnant models is constrained by the results of SPH, but its form is nonetheless
not crucial as long as significant mixing is not artificially introduced (via the
Vogt-Russell Theorem; Vogt 1926, Russell 1927). In comparing the predictions
of these models with the blue stragglers seen in two dense clusters, NGC 6397
and 47 Tuc, we have also assumed that all of the blue stragglers in these clusters
have been formed through collisions.
The apparent agreement between the predictions of our unequal mass merger
models and the observed blue straggler distributions NGC 6397 and 47 Tuc sug-
gests that
• little or no mixing occurred during the formation of these blue stragglers,
either of fresh hydrogen to the cores, or of helium to the surfaces. As
explained earlier, such mixing would produce a significant blueward scatter
in the distribution of blue stragglers in the CMD -- as it is, our models
with the best agreement with the observations, the unequal mass mergers,
produce rather red stars, relative to the clusters' ZAMS. Regardless of the
formation mechanism for the blue stragglers in these clusters, it is apparent
from their location redward of the ZAMS that they are formed as evolved
stars (Ouellette & Pritchet, 1996).
• there may be some form of efficient AML mechanism acting to slow the
rotation of these stars. Blue stragglers which are formed by collisions are
expected to be rapidly rotating: the affect of rapid rotation on the observed
distribution of blue stragglers in the CMD should result in a poor agree-
ment with our non-rotating models, and yet the agreement appears to be
quite good. The thin, short-lived convective envelopes seen in our models
precludes a magnetically driven stellar wind AML mechanism to explain the
slow rotation rates and apparent lack of rotational effects; however, angular
momentum transfer to a circumstellar disk or nearby companion are still
plausible.
The lack of surface convection in our most massive models and the thin con-
vective envelopes seen in our less massive models means that surface abundances
-- 24 --
of collisionally merged blue stragglers might not be altered from those of the
parent stars, unless some amount of hydrodynamical mixing occurs during the
collision. Additionally, meridional circulation currents, which should be confined
to the stellar envelope except in the most extreme cases, may act to mix the
surface layers, regardless of the amount of convection.
The extreme blue stragglers observed in both NGC 6397 and 47 Tuc occupy
the region of the CMD which should be populated by highly mixed stars. It is
possible that these blue stragglers are the few that were born rotating rapidly
enough that the meridional circulation currents are able to penetrate more deeply
in to the star; if so, these stars should still be rapidly rotating. It is, however,
possible that some fraction of these stars are merely photometric errors.
7. Acknowledgements
We would like to thank both Russell Robb and Don VandenBerg for providing
useful comments on an earlier version of this manuscript. Additional thanks must
go to Dr. VandenBerg for allowing us to use his stellar evolution code, and for
numerous insightful discussions on the topic of stellar evolution. Finally, we wish
to thank the Natural Sciences and Engineering Research Council for financial
support through a Research Grant to C.J.P..
-- 25 --
REFERENCES
Bailyn, C. D., 1995, Ann. Rev. Astron. & Astrophys. 33, 133
Bailyn, C. D. and Pinsonneault, M. H., 1995, Astrophys. J. 439, 705
Benz, W. and Hills, J. G., 1987, Astrophys. J. 323, 614
Benz, W. and Hills, J. G., 1992, Astrophys. J. 389, 546
Bolte, M., 1992, Astrophys. J., Suppl. Ser. 82, 145
Cameron, A. C., Campbell, C. G., and Quaintrell, H., 1995, Astron. Astrophys.
298, 133
Chaboyer, B., Demarque, P., and Sarajedini, A., 1996, Astrophys. J. 459, 558
Clement, M. J., 1994, Astrophys. J. 420, 797
Cox, J. P. and Giuli, R. T., 1968,
Gordon and Breach [1968]
in Principles of stellar structure, New York,
Edmonds, P. D., Gilliland, R. L., Guhathakurta, P., Petro, L. D., Saha, A., and
Shara, M. M., 1996, Astrophys. J. 468, 241
Glaspey, J. W., Pritchet, C. J., and Stetson, P. B., 1994, Astron. J. 108, 271
Guhathakurta, P., Yanny, B., Schneider, D. P., and Bahcall, J. N., 1992, Astron.
J. 104, 1790
Hills, J. G. and Day, C. A., 1976, Astrophys. Lett. 17, 87
Hobbs, L. M. and Mathieu, R. D., 1991, Publ. Astron. Soc. Pac. 103, 431
Hoffer, J. B., 1983, Astron. J. 88, 1420
Kaluzny, J., 1997, Astron. Astrophys. Suppl. Ser. 122, 1
Kaluzny, J., Kubiak, M., Szymanski, M., Udalski, A., Krzeminski, W., Mateo,
M., and Stanek, K., 1997, Astron. Astrophys. Suppl. Ser. 122, 471
Lang, K. R., 1992, in New York : Springer-Verlag, c1992-
Lauzeral, C., Ortolani, S., Auriere, M., and Melnick, J., 1992, Astron. Astrophys.
262, 63
Leonard, P. J. T., 1989, Astron. J. 98, 217
Leonard, P. J. T., 1996, Astrophys. J. 470, 521
Leonard, P. J. T. and Fahlman, G. G., 1991, Astron. J. 102, 994
Leonard, P. J. T. and Linnell, A. P., 1992, Astron. J. 103, 1928
Leonard, P. J. T. and Livio, M., 1995, Astrophys. J., Lett. 447, L121
-- 26 --
Livio, M., 1993, in R. A. Saffer (ed.), Blue Stragglers, Vol. 53 of ASP Conference
Series, p. 3
Lombardi, J. C. J., Rasio, F. A., and Shapiro, S. L., 1996, Astrophys. J. 468,
797 (LRS)
Mathieu, R. D. and Latham, D. W., 1986, Astron. J. 92, 1364
Mathys, G., 1991, Astron. Astrophys. 245, 467
Monaghan, J. J., 1992, Ann. Rev. Astron. & Astrophys. 30, 543
Mestel, L., 1953, Mon. Not. R. Astron. Soc. 113, 716
Ouellette, J. A. and Pritchet, C. J., 1996, in E. F. Milone and J.-C. Mermilliod
(eds.), The Origins, Evolution, and Destinies of Binary Stars in Clusters.,
Vol. 90 of ASP Conference Series, p. 356
Ouellette, J. A. and Pritchet, C. J., 1998b, In preparation
Paresce, F., Meylan, G., Shara, M., Baxter, D., and Greenfield, P., 1991, Nature
352, 297
Peterson, R. C., Carney, B. W., and Latham, D. W., 1984, Astrophys. J. 279,
237
Podsiadlowski, Ph., 1996, Mon. Not. R. Astron. Soc. 279, 1104
Pritchet, C. J. and Glaspey, J. W., 1991, Astrophys. J. 373, 105
Rodgers, A. W. and Roberts, W. H., 1995, Astron. J. 109, 264
Russell, H. N., 1927, in Russell, Dugan, and Stewart (eds.), Astronomy, Vol. 2,
p. 910
Sandquist, E. L., Bolte, M., and Hernquist, L., 1997, Astrophys. J. 477, 335
Sarajedini, A. and Forrester, W. L., 1995, Astron. J. 109, 1112
Schwarzschild, M., 1958,
in Structure and evolution of the stars., Princeton,
Princeton University Press
Shara, M. M., Saffer, R. A., and Livio, M., 1997, Preprint
Shore, Steven N., 1992,
in An Introduction to Astrophysical Hydrodynamics,
Academic Press [1992]
Sigurdsson, S. and Phinney, E. S., 1993, Astrophys. J. 415, 631
Sills, A., Bailyn, C. D., and Demarque, P., 1995, Astrophys. J., Lett. 455, L163
Sills, A., Lombardi, J., Bailyn, C., Demarque, P., Rasio, F., and Shapiro, S.,
1997, Preprint
-- 27 --
Sills, A. and Lombardi, James C., J., 1997, Astrophys. J., Lett. 484, L51
Stryker, L. L., 1993, Publ. Astron. Soc. Pac. 105, 1080
Tassoul, M. and Tassoul, J. L., 1984, Astrophys. J. 279, 384
Tout, C. A. and Pringle, J. E., 1992, Mon. Not. R. Astron. Soc. 256, 269
VandenBerg, D. A., Swenson, F. J., Rogers, F. J., Iglesias, C. A., and Alexander,
D. R., 1997, In preparation
Vogt, H., 1926, Aston. Nachr. 226, 301
This preprint was prepared with the AAS LATEX macros v4.0.
-- 28 --
Fig. 1. -- Density (ρ) and entropy (S) profiles for representative polytropes.
Shown are profiles for 0.80M⊙ (solid line, n = 3, R = 1.0R⊙), 0.60M⊙ (dashed
line, n = 3, R = 0.56R⊙), and 0.40M⊙ (dotted line, n = 3/2, R = 0.37R⊙)
polytropes, calculated using the radii and polytropic indices given. The 0.6M⊙
composite polytrope has an n = 3 core and an n = 1.5 envelope with the bound-
ary between the two located at Rbnd = 0.29R⊙.
Fig. 2. -- Density (ρ) and entropy (S) profiles for real GC stars. Shown are
profiles for 0.864M⊙ (solid line), 0.60M⊙ (dashed line), and 0.40M⊙ (dotted line)
stellar models at an age (14 Gyr) and metallicity ([Fe/H]=-0.71) appropriate for
47 Tuc.
Fig. 3. -- Comparison of pre-shock (thin lines) and post-shock (thick lines) en-
tropy profiles for the stars shown in Figure 2. See Section 3.1.2 for a description
of the calculation of the amount of entropy production during shock heating
Fig. 4. -- Evolutionary tracks (thin solid lines) for equal mass (top) and unequal
mass (bottom) merger models for NGC 6397. Masses of the models range from
0.90M⊙ to 1.40M⊙ in steps of 0.10M⊙. The open circles (◦) are placed along the
tracks at equal intervals of 0.05Gyr. Also shown are the theoretical cluster ZAMS
(thick solid line) and a 19 Gyr isochrone (thick dashed line). Note the lack of a
convective "hook" in all of the models, except for the most massive equal mass
merger model.
Fig. 5. -- Similar to Figure 4, except for 47 Tuc. The masses of the models range
from 1.10M⊙ to 1.70M⊙ in steps of 0.10M⊙. Also shown is a 14 Gyr isochrone.
Symbols and line styles have the same meanings as in Figure 5.
Fig. 6. -- The observed distribution of blue stragglers in NGC 6397 (•) compared
with blue stragglers created from the tracks of our merger models (✷, see text).
Also shown are the tracks for the appropriate mergers with the 'pre-blue straggler'
phase omitted for clarity (thin dotted lines). The error bar at the top of the figure
corresponds to the random photometric errors appropriate for the observations.
The theoretical cluster ZAMS (thick solid line) and a 19Gyr isochrone (thick
dashed line) are also shown.
Fig. 7. -- Similar to Figure 6, except for 47 Tuc. Also shown is a 14 Gyr isochrone.
Symbols and line styles have the same meanings as in Figure 6.
-- 29 --
Table 1: Cluster Parameters
Parameter NGC 6397
47 Tuc
Accepted Parameters
[Fe/H]
[α/Fe]
E(B − V )
V(HB)
ca
-1.91
0.30
0.18
12.90
2.50
Derived Parameters
V
M HB
(m − M)V
Age(Gyr)
MT O(M⊙)
0.598
12.30
19.0
0.708
-0.71
0.30
0.04
14.09
2.04
0.838
13.25
14.0
0.864
a Central concentration. Defined as c = log(rtidal/rcore).
-- 30 --
-- 31 --
-- 32 --
-- 33 --
-- 34 --
-- 35 --
-- 36 --
|
astro-ph/0302219 | 2 | 0302 | 2004-07-09T19:16:57 | The Effect of Neutron Star Binding Energy on Gravitational-Radiation-Driven Mass-Transfer Binaries | [
"astro-ph"
] | In a relativistic model of a neutron star, the star's mass is less than the mass of the individual component baryons. This is due to the fact that the star's negative binding energy makes a contribution to the star's total energy and its mass. A consequence of this relativistic mass deficit is that a neutron star that is accreting matter increases its mass at a rate which is slower than the mass of a baryon times the rate that baryons are accreted. This difference in the rate of change of the masses has a simple relation with the star's gravitational redshift. We show that this effect has the potential to be observed in binaries where the mass transfer is driven by angular momentum losses from the gravitational radiation emitted by the binary motion. | astro-ph | astro-ph |
Accepted by the Astrophysical Journal, to appear Oct 2004.
Preprint typeset using LATEX style emulateapj v. 11/12/01
THE EFFECT OF NEUTRON STAR GRAVITATIONAL-BINDING ENERGY ON
GRAVITATIONAL-RADIATION-DRIVEN MASS-TRANSFER BINARIES
Theoretical Physics Institute, Department of Physics, University of Alberta, Edmonton AB, Canada,
Evelyne Al´ecian1 and Sharon M. Morsink
Accepted by the Astrophysical Journal, to appear Oct 2004.
T6G 2J1
ABSTRACT
In a relativistic model of a neutron star, the star's mass is less than the mass of the individual
component baryons. This is due to the fact that the star's negative binding energy makes a contribution
to the star's total energy and its mass. A consequence of this relativistic mass deficit is that a neutron
star that is accreting matter increases its mass at a rate which is slower than the mass of a baryon
times the rate that baryons are accreted. This difference in the rate of change of the masses has a simple
relation with the star's gravitational redshift. We show that this effect has the potential to be observed in
binaries where the mass transfer is driven by angular momentum losses from the gravitational radiation
emitted by the binary motion, if the physics of the donor star is well-enough understood.
Subject headings: stars: neutron -- relativity -- gravitational radiation -- binaries -- accretion
1.
introduction
In the general theory of relativity, all forms of stress
and energy act as a source for the gravitational field. As
a result, the mass of a compact star is the total energy of
the star divided by c2. As the total energy of a star in-
cludes both the rest-mass energy and the negative binding
energy, the mass of the star is smaller than the sum of the
masses of its component particles. This mass deficit effect
is roughly proportional to the star's compactness, so it is
most important for neutron stars. For neutron stars near
the upper mass limit, the mass deficit can be as large as
25% of the star's mass (see for example, Cook, Shapiro &
Teukolsky (1994)), depending on the equation of state.
A consequence of the mass deficit is that when a neu-
tron star accretes matter, its mass changes at a rate that
differs from the rate that baryons are accreted times the
mass of a baryon. This mass deficit effect is then poten-
tially observable through changes in the orbital period of
the binary caused by accretion, if the mass accretion rate
is known. An important class of binaries where this effect
may be observable are binaries in which the mass transfer
is driven by angular momentum losses due to the emission
of gravitational radiation caused by the orbital motion.
In the standard scenario for gravitational-radiation-driven
mass transfer (Kraft et al. 1962; Faulkner 1971), the mass
deficit is neglected since the primaries are white dwarf
stars. However, in binaries where the primary is a neu-
tron star, the effect is in principle observable, and could
lead to constraints on the neutron star mass and equa-
tion of state. In this paper we will address this issue and
discuss the conditions where the relativistic mass deficit's
effects could be detected. The effect described requires
very precise timing of the binary's orbital period in order
to be detected, so we will focus our attention on the class
of binaries containing an accreting millisecond X-ray pul-
sar. At present three binaries of this class have measured
mass functions: SAX J1808.4-3658 (Wijnands & van der
Klis 1998; Chakrabarty & Morgan 1998), XTE J1751-305
(Markwardt et al. 2002) and XTE J0929-314 (Galloway
et al. 2002). In addition two more pulsars, XTE J1807-294
(Markwardt et al 2003) and XTE J1814-338 (Strohmayer
et al 2003) have recently been discovered, although their
mass functions are not yet measured. Precise timing of the
X-ray pulse arrival times allows the detection of these bi-
nary systems' orbital periods through the orbital Doppler
effect. It is expected (Chakrabarty & Morgan 1998) that
future observations will reveal an orbital period derivative.
The effect of changing a neutron star's binding energy
in a binary system has been examined by other authors in
other situations. In the case of a binary system without
mass-transfer, Spyrou & Stergioulas (2001) considered the
changes in a neutron star's binding energy due to mag-
netic dipole spin-down and its effect on the orbital pe-
riod. The magnitude of the effect examined by Spyrou &
Stergioulas (2001) is quite small, but may be measurable.
In the present paper we allow for mass transfer, which
produces an effect that is much larger. While accretion-
induced changes in a neutron star's characteristics have
been considered by Spyrou (1988), his model was for a sim-
ple Newtonian stellar model and did not consider changes
in the orbital period.
In section 2 we review the concept of the binding energy
of a star and discuss its value and its derivatives for rel-
ativistic models of neutron stars. In section 3 we review
the scenario of gravitational-radiation-driven mass trans-
fer and include the mass deficit in the derivation. In sec-
tion 4 we review the different theories of how the compact
binaries of interest may have evolved and discuss which
theories would allow the desired measurements of the rel-
ativistic effect. Finally in section 5 we discuss the main
theoretical uncertainties in this method.
2. the binding energy of compact objects
If we define N to be the total number of baryons in the
star, mB the mass of a baryon and ε the binding energy
of the star, then the total mass, M , of the star is
(1)
The binding energy is the sum of the star's gravitational
potential energy, the internal energy and the rotational
M = mBN + ε/c2.
1 Present Address: Universit´e Paris VII, 2 place Jussieu, 75251 Paris cedex 05, France
1
2
kinetic energy.
If the star were to be disassembled into
its component baryons, the mass of the baryons would be
M0 = mBN . For a gravitationally bound system, the
binding energy is negative, so M is always smaller than
M0. Using Newtonian arguments, the leading order con-
tribution to the binding energy is ε ∼ −GM 2
0 /R, so that
the mass is given by
M = M0(cid:18)1 − ε0
GM0
c2R (cid:19)
(2)
where ε0 is a constant of order unity. Equation (2) shows
that the relativistic correction to the mass of a star scales
as the compactness GM/(Rc2). Hence this correction is
negligible for main sequence stars and white dwarfs. How-
ever, for neutron stars, the compactness can be as large as
4/9. The contribution of the binding energy to the mass
can be significant for compact stars.
It is useful to briefly review the concept of mass and
binding energy by recalling their definitions for a spherical
star in general relativity (see box 23.1 of Misner, Thorne
& Wheeler (1973) for an elementary discussion). A zero
temperature equation of state (EOS) can be written as
equations for P (n) and ǫ(n) where n is the number density
of baryons, P is the pressure and ǫ is the energy density.
The energy density includes both the matter's rest-mass
energy density and its internal energy density. Once the
EOS is specified, the star's total mass (using Schwarzschild
coordinates) is
M = 4πZ R
0
ǫr2dr.
(3)
This is the mass which is measured by Kepler's law
when a satellite orbits the star. For this reason, M
is often called the "gravitational mass". The baryon
mass, M0, of the star is given by the volume integral
of the baryon number density times mB.
Since the
volume element in Schwarzschild coordinates is dV =
4πr2 (1 − 2m(r)/r)−1/2 dr, the baryon mass is
M0 = mBN = 4πmBZ R
n (1 − 2m(r)/r)−1/2 r2dr.
Even in the case of an incompressible fluid, there is a dif-
ference between M and M0 due to the binding energy of
the star. These definitions can be extended to rotating
stars (see for example, Friedman et al. (1986)). The result
is a family of equilibrium stellar models with gravitational
masses that depend on the star's baryon number N and
its angular momentum J (Bardeen 1973).
(4)
0
Consider the accretion of baryons with specific angular
momentum ℓ at the rate N onto a star. Since the grav-
itational mass is a function of the stars' baryon number
and angular momentum, the rate of increase of the gravi-
tational mass is
∂N (cid:19)J
M = N(cid:18)(cid:18) ∂M
+ mBℓ(cid:18) ∂M
∂J (cid:19)N(cid:19) .
The partial derivatives of the gravitational mass are given
by (Bardeen 1973)
where Ω is the star's angular velocity, and
(cid:18) ∂M
∂J (cid:19)N
(cid:18) ∂M
∂N (cid:19)J
=
Ω
c2 ,
= mBΦ,
(5)
(6)
(7)
where Φ is the dimensionless chemical potential (or injec-
tion energy) for bringing one baryon from infinity to the
pole of the rotating star. Using the metric of an axisym-
metric rotating star, given by Friedman et al. (1986)
ds2 = −e2νdt2 + e2ψ (dφ − ωdt)2 + e2µ(cid:0)dr2 + r2dθ2(cid:1) , (8)
the chemical potential is given by
Φ = (eν)p ,
(9)
where the subscript "p" denotes that the function is eval-
uated at the star's spin pole. In the limit of a nonrotating
star,
lim
Ω→0
Φ =r1 −
2GM
Rc2 .
(10)
Typical values for the chemical potential for rotating neu-
tron stars can be found in the paper by Friedman et al.
(1986), however they define a quantity β that is related to
the chemical potential by Φ = √β. The chemical potential
can be thought of as the specific energy required to bring
a particle from infinity to the star. Since a particle at rest
at infinity has larger energy than a particle at rest at the
surface of the star (due to gravitational redshift), energy
is liberated during accretion. The quantity 1 − Φ is the
fraction of the energy liberated during accretion.
If we define a dimensionless angular velocity, ¯Ω
¯Ω = Ωr R3
GM
and a dimensionless specific angular momentum, ¯ℓ,
¯ℓ =
cℓ
M G
,
(11)
(12)
the rate of change of the star's gravitational mass due to
accretion is
M = mB N Φ + ¯ℓ ¯Ω(cid:18) GM
c2R(cid:19)3/2! .
(13)
In general, not all of a particle's angular momentum will
be available to be added to the star's angular momentum.
External torques, such as the star's magnetic field (Ghosh
& Lamb 1991) or gravitational radiation (Wagoner 1984;
Andersson et al. 1999) from nonaxisymmetric perturba-
tions may remove the angular momentum.
It is useful
then to introduce a torque parameter η defined so that if
the baryon is in a circular orbit at radius r around the star
when it is accreted, the specific angular momentum added
to the star is
¯ℓ = η ¯ℓK,
(14)
where ¯ℓK is the dimensionless specific angular momentum
of a particle in a circular Keplerian orbit. In the limit of
a nonrotating star,
¯ℓK =
lim
Ω→0
rc2
GMs
M
rc2/G − 3M
.
(15)
For a rotating star ¯ℓK is given by the quantity cL/M G,
where L is the specific angular momentum defined in equa-
tion (7) of Cook, Shapiro & Teukolsky (1994). Since the
rate that the star's mass increases will be close to mB
APR
C
L
B60
B80
3
1.5
2
2.5
3
Mass (Solar Mass Units)
γ
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
1
Fig. 1. -- Plots of the parameter γ versus neutron star mass for different models of nonrotating neutron stars.
multiplied by the rate that baryons are accreted, we will
introduce a parameter γ, defined by
γ = 1 −
M
mB N
= 1 − Φ − η ¯ℓK ¯Ω(cid:18) GM
c2R(cid:19)3/2
,
(16)
which is roughly proportional to the star's compactness.
The parameter γ corresponds to the fractional rate of
change of the relativistic mass deficit. For a slowly ro-
tating star,
lim
Ω→0
γ = 1 −r1 −
2GM
Rc2 .
(17)
Recall that for a non-rotating star, the gravitational red-
shift, z, is defined by
z =
1
q1 − 2GM
Rc2
− 1,
(18)
so in the limit of spherical symmetry, the parameter γ
is related to z by γ = z/(z + 1). For rotating stars,
this relationship can be generalised by defining zp as
the gravitational redshift from the star's pole. The pa-
rameter γ is then related to the polar redshift by γ =
.
zp/(zp + 1) − η ¯ℓK ¯Ω(cid:0)GM/Rc2(cid:1)3/2
In Figure 1 plots of γ versus gravitational mass for non-
rotating neutron stars are shown for a few representative
equations of state. We have chose three traditional neu-
tron star equations of state (labeled C, L and APR) and
two simplified quark star (labeled B60 and B80) equations
of state. From the Arnett & Bowers (1977) catalogue, we
have chosen equation of state C (Bethe & Johnson 1974)
which is moderately stiff and L (Pandharipande & Smith
1975) which is very stiff in order to show a typical range
of stiffness for equations of state. Equation of state APR
(Akmal et al. 1998) makes use of data from modern nu-
cleon scattering experiments and also includes first order
special relativistic corrections. The quark star equations
of state are constructed in an approximation described in
Glendenning (2000) where the quarks' masses are set to
zero and the MIT bag model is used. The number after
the letter "B" in the quark equations of state refer to the
value of the bag constant, B, in units of MeV/fm3.
The effects of rotation are shown in Figure 2, where
values of γ for the APR equation of state are plotted for
different spin frequencies νs and torque values. The rela-
tivistic parameter γ was computed by evaluating equation
(16) after computing the value of the rotating star metric
given in equation (8) using an accurate 2D code1 written
by Stergioulas and described in Stergioulas & Friedman
(1995). This code is, in turn, based on the methods de-
scribed by Komatsu, Eriguchi & Hachisu (1989) and Cook,
Shapiro & Teukolsky (1994).
In the case where the ac-
creted matter does not add any angular momentum to the
star (η = 0), it can be seen in Figure 2 that the func-
tion γ is almost independent of the star's spin frequency.
The plots labeled η = 1 correspond to stars where the
magnetic field's effect on the accretion disk is neglected
so that the orbits extend down to the marginally stable
circular orbit and the accreted matter adds all of its an-
gular momentum to the star.
In this case, the effect of
the star's spin is to decrease γ by a large fraction over
the nonrotating case. Since the value of ℓK has very little
angular velocity dependence, it can be seen that the ef-
fect of the star's angular velocity is to approximately shift
the curve downwards by an amount linear in the angu-
lar velocity.
If the torque parameter were negative, the
curves would be shifted vertically upwards by an amount
linear in the angular velocity. The magnetic disk models
of Rappaport, Fregeau, & Spruit (2004) allow material to
be accreted without being propellered away from the star
1 The code is available at: http://www.gravity.phys.uwm.edu/rns/
γ
0.55
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
1
s = 400 Hz, η = -1
ν
s = 200 Hz, η = -1
ν
ν
s = 0 Hz
ν
s = 200 Hz, η = 0
s = 400 Hz, η = 0
ν
s = 200 Hz, η = 1
ν
ν
s = 400 Hz, η = 1
1.5
2
2.5
3
Mass (Solar Mass Units)
Fig. 2. -- Plots of the parameter γ versus neutron star mass for equation of state APR, for different stellar spin frequencies (νs) and torque
parameters (η).
4
while decreasing the star's angular momentum. In their
models, the torque parameter can be negative and have
a magnitude which is of order unity. The curves labelled
η = −1 correspond to this scenario and have an increased
value of γ.
Simultaneous measurements of a neutron star's mass
and the parameter γ have the potential to constrain the
neutron star equation of state. Since γ is defined through
M and mB N , accurate measurements of
the difference of
these derivatives would be desirable. Clearly these are not
trivial measurements. In the highly idealized case of mass
transfer driven by gravitational radiation, there is the po-
tential to make this measurement if enough is known about
the companion star.
In the next section we review the
gravitational-radiation-driven mass-transfer scenario, and
discuss how γ could be measured in an ideal situation.
We will also discuss the uncertainties in the physics of the
scenario that will likely limit the ability to measure γ.
3. gravitational-radiation-driven mass transfer
We now review the basic scenario for mass transfer
driven by angular momentum losses due to gravitational
radiation. The gravitational-radiation-driven mass trans-
fer scenario was first developed (Kraft et al. 1962; Faulkner
1971) to explain the properties of white dwarf binaries with
short periods, however the physics also applies to neutron
star binaries. In this section we will give a short review
of the derivation of the equations including the effects of
both mass loss and the neutron star's binding energy. The
derivation of these equations excluding the binding energy
effect appears in many sources, such as Rappaport et al.
(1982) and Verbunt (1993). We will make use of the fol-
lowing notation: M1 is the neutron star's gravitational
mass, while M2 and R2 are the mass and radius of the
companion star. The mass ratio is denoted q = M2/M1.
For the ultra-compact systems which are most likely to be
of interest, the mass ratio is probably very small.
If there is no mass transfer between the stars, the or-
bital separation will slowly shrink due to the emission of
gravitational radiation. The binary system loses angular
momentum at the rate
J!GR
J
64π
5
= −
q(1 + q)−1/3(cid:18) 2πGM1
c3P (cid:19)5/3 1
P
(19)
where P is the orbital period. As the stars move closer to-
gether the Roche lobe radius of the secondary star shrinks
and eventually meets the surface of the secondary, at which
point mass will begin to be transfered to the neutron star.
Once mass transfer begins, the size of the secondary star
is assumed to be the same as the Roche lobe (RL).
In
the Paczy´nski (1971) approximation, the size of the Roche
lobe (and R2) is given by
R2 = RL = 0.46a(cid:18) q
1 + q(cid:19)1/3
,
(20)
where a is the distance between the stars. The time deriva-
tive of R2 is
R2
R2
=
a
a
+
1
3
q
q
1
(1 + q)
.
(21)
a white or brown dwarf), we do not need to make a dis-
tinction between the gravitational and baryon masses of
the secondary. As the secondary loses mass, its radius is
assumed to change as
R2
R2
= nad
M2
M2
+ R2
R2!th
,
(22)
where nad describes the purely adiabatic changes to the
star that depend only on the mass loss, while (cid:16) R2/R2(cid:17)th
corresponds to changes in the star's structure due to ther-
mal adjustments. This last term is of order 1/τKH, where
τKH is the Kelvin-Helmholtz timescale. While in most
cases the thermal timescale can't be neglected, in the case
of a cool white dwarf the Kelvin-Helmholtz timescale will
be long enough that only terms depending on the mass
loss rate need to be included in equation (22).
In the most naive model of a degenerate star, nad =
−1/3. However for very low mass white dwarfs with finite
temperature and an equation of state including Coulomb
repulsion, the values of nad range from -1/3 to 0 (Deloye &
Bildsten 2003), depending on the mass, composition and
temperature. If the mass and composition are known, the
uncertainty in temperature introduces an uncertainty in
nad of about 10%.
For now, we will allow the most general mass transfer
and assume that only a fraction β of the mass lost from
the secondary is captured by the neutron star. As a result,
the baryon mass of the neutron star increases at the rate
(23)
However, the gravitational mass of the neutron star in-
creases at a slower rate, given by
M0,1 = −β M2.
M1 = −(1 − γ)β M2,
(24)
where the parameter γ (discussed in the previous section)
vanishes for newtonian stars and is approximately equal
to the neutron star's compactness.
Since we are allowing for non-conservative mass trans-
fer, angular momentum will be lost. Following the work of
Rappaport et al. (1982), we parametrize this angular mo-
mentum loss due to non-conservative mass transfer with
the parameter α, so that
= α(1 − β)
2πa2
P
M2.
(25)
(cid:16) J(cid:17) M
The parameter α encompasses the uncertainty in where
the mass is lost. If the mass loss occurs close to the centre
of mass (say near the neutron star) then α will be very
small (of order q2, see Podsiadlowski, Rappaport, & Pfahl
(2002)), while if the mass loss occurs near the low mass
secondary, α will be close to unity.
The liberation of energy during accretion due to the ef-
fects discussed in this paper also introduce a mechanism
for angular momentum loss. During the accretion process,
the number of baryons and angular momentum transfered
to the star are conserved, while energy is liberated at the
rate E = βγ M2c2 in the rest frame of the star. How-
ever, the star is orbiting the centre of mass at a distance
of a1 = qa/(1 + q). Introducing a parameter δ, of order
unity, the angular momentum loss due to the orbital mo-
tion of the neutron star as the energy is liberated is
M2. Since the sec-
The secondary loses mass at the rate
ondary is assumed to be a non-relativistic star (such as
= δβγ
(cid:16) J(cid:17)γ
2πa2
P (cid:18) q
1 + q(cid:19)2
M2.
(26)
In the systems of interest, the mass ratio q is likely to be
very small, so this term will not be of much importance.
Another mechanism for angular momentum loss is mag-
netic braking, where the donor star's stellar wind causes
the donor to lose angular momentum through a coupling
with the star's magnetic field.
If the binary system is
tidally locked to the donor's spin, this causes the binary
system to lose orbital angular momentum (Verbunt &
Zwaan 1981). We denote this mechanism through a term
JB which is independent of the mass loss rate. Estimates
of the magnetic braking law depend on knowledge of main
sequence stellar winds and rotation rates. This type of
mechanism is most likely to be of importance if the donor
was originally a main sequence star. The effects of different
parameterizations of JB have been explored by Rappaport,
Verbunt, & Joss (1983).
Since the mass accreted onto the neutron star adds an-
gular momentum to the star, this too creates an effective
loss of orbital angular momentum. However, since this loss
is proportional to (R1/a)2, (where R1 is the neutron star's
radius) it is negligible.
Combining together all the different mechanisms for loss
of orbital angular momentum, the total angular momen-
tum loss rate is
J
J
= J
J!GR
+ J
J! M
+ J
J!γ
J!B
+ J
.
Since the binary's orbital angular momentum is
(27)
(28)
J = M1M2(cid:18) Ga
M1 + M2(cid:19)1/2
the change in orbital angular momentum due to rearrange-
ments of the mass is
J! =
J
1
2
a
a
+
M1
M1
+
M2
M2 −
1
2
( M1 + M2)
(M1 + M2)
.
(29)
Equations (21), (22), (24), (27) and (29) are a set of equa-
tions that allows the mass accretion rate to be written in
terms of the angular momentum loss rate due to various
processes,
M2
M2
A(α, β, γ, nad, q)" J
2 R2
+ J
J!B
J!GR
−
3
2
=
1
1
R2!th#
(30)
where the function A(α, β, γ, nad, q) is defined by
A =
5 + 3nad
4
3
2
−
q(1 − γ)β −
1
2
3
2
−
α(1 − β)(1 + q) −
3
2
δβγ
q
1 + q
q2
1 + q
,
(1 − β + βγ)
(31)
and it should be remembered that the adiabatic index nad
depends on M2 = qM1. When γ = 0, this equation re-
duces to the equation derived by Rappaport et al. (1982)
for non-conservative mass transfer. In the binary systems
of interest q is small, so the term proportional to δ is ex-
plicitly very small. It should also be remembered that the
constant α is proportional to q2 if the mass loss occurs near
to the neutron star, which is the most likely situation.
If we were to ignore changes in the secondary due to
thermal adjustments and any angular momentum losses
due to magnetic braking, a formula for the orbital period
derivative can be derived. (We will the discuss the appli-
cability of these assumptions in the next section.) Making
use of Kepler's law, we find the following equation for the
orbital period derivative
5
P =
96π
5
q(1 + q)−1/3
B(α, β, γ, nad, q)(cid:18) 2πGM1
c3P (cid:19)5/3
,
(32)
where the function B(α, β, γ, nad, q) is defined by
B =
2
A.
1 − 3nad
Making use of the mass function f ,
(q sin i)3
(1 + q)2 ,
f = M1
(33)
(34)
the orbital period derivative is
P =
96π
5 (cid:18) f (1 + q)
M⊙ (cid:19)1/3
1
B sin i(cid:18) 2πGM⊙
c3P (cid:19)5/3(cid:18) M1
M⊙(cid:19)4/3
(35)
.
The predicted values of
P can be written in the form
P = D
B × 10−14 (1 + q)1/3
sin i
M⊙(cid:19)4/3
(cid:18) M1
,
(36)
where the observationally determined constant D is de-
fined by
D = 18.40 ×
f 1/3
−6
P 5/3
3
,
(37)
and f /M⊙ = f−6×10−6 and P = P3×103 s. Values for the
constant D for the ultra-compact binaries with measured
mass functions are given in Table 1. Since the contribu-
tion to the period derivative from the relativistic term γ
is quite small, we will typically require 3 significant fig-
ures, hence the displayed accuracy in equation (37) and in
Table 1. Although the binary mass ratio q is quite small
for the systems of interest we have explicitly kept the fac-
tor of (1 + q)1/3 in the equation for the period derivative
since a precise result is required. Present observations of
the orbital period are not precise enough to detect such a
small period derivative in the orbit of any of the systems
listed in Table 1, but observations over a long period of
time may allow the detection of the period derivative.
A similar formula for the rate that mass is lost from the
companion can be written in the form
A(sin i)2 (cid:18) M1
M2 = F × 10−12 (1 + q)
where the constant F is defined by
f 2/3
−6
P 8/3
3
F = 58.1 ×
M⊙(cid:19)2 M⊙
yr
,
(38)
.
(39)
It has
Values for the constant F are given in Table 1.
already been shown by Chakrabarty & Morgan (1998);
Bildsten & Chakrabarty (2001) that the mass transfer rate
for SAX J1808.4-3658 is consistent with being driven by
M2
gravitational radiation. However, the determination of
is probably no better than order of magnitude, so these
measurements can't be thought of as providing an extra
constraint on γ.
6
4. astrophysical considerations
The expression for
the period derivative derived
in the previous section depends on the parameters
M1, sin i, nad, α, β, δ and γ. (The mass ratio q is not inde-
pendent, since it is determined once f is measured and M1
and sin i are specified.) This expression was derived by as-
suming that the secondary's thermal timescale is very long,
and that magnetic braking is not operating. We now con-
sider some of the models for the evolution of ultra-compact
binaries in order to evaluate in which models there might
be a chance of detecting γ.
There are two main scenarios for producing a neutron
star X-ray binary with a very short orbital period (see
Nelson & Rappaport (2003) for a discussion). In the first
scenario, the companion star is a main sequence star and
will be denoted MS-NS. The alternative picture is of a
white dwarf companion and will be denoted WD-NS.
In the MS-NS scenario, mass transfer from a low mass
star occurs near the time when Hydrogen is depleted in its
core (Nelson & Rappaport 2003). Both gravitational radi-
ation and magnetic braking (Ergma & Antipova 1999) re-
move orbital angular momentum from the system, and the
orbital periods can evolve down to very short timescales
(Podsiadlowski, Rappaport, & Pfahl 2002). In this type
of system, magnetic braking could easily be as important
as gravitational radiation in the removal of angular mo-
mentum. Since the parameters describing magnetic brak-
ing are not accurately known, it would not be possible to
measure the effect of the γ term in this type of binary
system.
In the WD-NS scenario the white dwarf fills its Roche
lobe when the orbital period is very small (3 minutes for
example) and gravitational radiation drives mass transfer
and increases the orbital period to the range of 40 min-
utes, reducing the white dwarf mass to very low values of
order 0.01M⊙ Rasio, Pfahl, & Rappaport (2000). Mag-
netic braking is unlikely to be operating in the WD-NS
picture, since it is most effective if the donor star has a
large radius, and it typically assumes a convective enve-
lope that drive a magnetic dynamo. A more difficult ques-
tion is whether the white dwarf is being heated, through
either being irradiated by the X-ray flux from the neutron
star or through tidal effects.
If the heating timescale is
short compared to the mass-loss timescale, the ( R2/R2)th
term can't be discarded. Deloye & Bildsten (2003) have
M2 and P for WD-NS systems
shown that observations of
can constrain possible combinations of WD composition
and temperature and have demonstrated that it is possi-
ble to use these observations to show that in some cases
an adiabatic evolution is consistent with observations.
If a measurement of the orbital period derivative of a
WD-NS binary were to be made, and if one assumed that
the evolution is adiabatic, equation (35) could be used to
constrain a combination of M1, sin i, nad, α, β, δ and γ. In
the compact systems most likely of interest, q is proba-
bly quite small, of order 0.01, so terms of order q2 in the
function A (such as the term proportional to δ) can be ne-
glected. If any mass loss take place near the neutron star
(say through a propeller mechanism) then the parameter
α ∼ q2 and the α term can be neglected. In the function
A, γ always appears in the combination qβ(1 − γ), so it is
essentially this term that we would like to measure. Since
γ could realistically be expected to have a value near 0.2
if we were pessimistic and supposed that β could be as
small as 0.5, the term qβ(1 − γ) ≥ 4 × 10−3. We would
then want to know nad to an accuracy of at least ±0.001
in order to have a hope of measuring the combination of
β(1 − γ), which would give an upper limit on the value of
γ.
If the WD temperature and composition were known
then nad is given by Deloye & Bildsten (2003) as a function
of M1 and sin i. For a given WD temperature and compo-
sition, a period derivative measurement (along with all of
the other assumptions listed in this section) would yield a
constraint on a combination of M1, sin i, and β(1−γ). Fur-
ther constraints on the inclination angle would strengthen
this constraint. For example recent observations of SAX
J1808.4-3658 by Wang et al. (2001) have made use of a
reddening distance technique and models of X-ray heated
accretion disks to constrain the binary inclination, depend-
ing on the assumed distance to the binary system.
In the equations for the period derivative for mass trans-
fer induced only by gravitational radiation with adiabatic
changes to the white dwarf, the value of the nad plays the
most important role (along with the inclination angle and
the neutron star mass) in determining the binary's evo-
lution.
Ideally, observations of such as system ignoring
the relativistic γ term could be used to constrain ranges
of values of nad consistent with an adiabatic evolution for
various possible values of sin i and M1. Once these values
are determined, and expanding equation (31) to lowest
order in q (assuming that the mass loss occurs near the
neutron star) one finds the limit on γ,
3
2 −
1
q (cid:18) (5 + 3nad)
(1 − 3nad)
4
D × 10−14
2
P
γ ≤
−
sin i
(1 + q)1/3
M⊙(cid:19)4/3!(40)
(cid:18) M1
where we have used the fact that β ≤ 1. Although a num-
ber of assumptions have been made in deriving this limit,
inequality (40) does have the potential to provide an in-
teresting limit on γ. Since the mass ratio q is expected to
be small (of order 0.01), a meaningful limit will only occur
if the terms proportional to 1/q are similar in magnitude,
so that their difference is of order q. Since D is of order
unity, only orbital period derivatives of order 10−14s/s will
provide an interesting limit on γ, given all of our assump-
P is too
tions. From equation (40) it can be seen that if
small, the maximum allowed value of γ will be so large as
P is too large,
to be an uninteresting limit. If however,
the maximum allowed value of γ will be negative, implying
that our assumptions about the physics of the binary are
invalid.
As an example, consider the situation where the com-
panion star is a Helium white dwarf with mass 0.02M⊙
and an adiabatic index of nad = −0.22, consistent with a
temperature near 107K as calculated in the models of De-
loye & Bildsten (2003). With data for the mass function
for XTE J1751-305, inclination angles with sin i ≥ 0.7 give
neutron star masses larger than 1.4 M⊙. Given this range
of inclination angles, it is easy to use equation (40) to find
the maximum allowed value of γ for any observed P . These
values are shown in Figure 3. This figure can be inter-
preted by noting that if the binary's inclination angle were
2.9
3.1
3.3
3.5
3.7
3.9
4.1
1
0.8
0.6
0.4
0.2
γ
m
u
m
x
a
M
i
7
Fig. 3. -- Plots of the maximum allowed values of γ versus binary inclination angle for XTE J1751-305, assuming a Helium white dwarf
companion with M2 = 0.02M⊙ and an adiabatic index of nad = −0.22. In this plot, each curve corresponds to a value of the orbital period
derivative in units of 10−14s/s.
0
0.6
0.65
0.7
0.75
0.8
sin(i)
0.85
0.9
0.95
1
P = 3.9× 10−14s/s
found to be sin i = 0.9, then a value of
would be consistent with our assumptions, while a larger
period derivative (say 4.1 × 10−14) would yield a negative
γ and is inconsistent with our assumptions.
5. conclusions
In this paper we have discussed the effect on the orbital
period of changing a neutron star's binding energy through
accretion. We have shown that it is possible, in principle,
for an observation of an orbital period derivative and with
some knowledge of the donor star to place constraints on
possible values of the neutron star's mass and the binding
energy parameter that depend on the binary's inclination
angle. However, there are a number of uncertainties in-
herent in our method. We will now discuss these potential
problems.
Our calculation assumes that the binary's orbit is given
by newtonian physics and have not considered post-
newtonian (PN) corrections. However, the 1st PN cor-
rections enter at the level of v2/c2 where v is the orbital
velocity. Since we are considering wide (from a relativist's
point of view) binaries with separations of the order of
104 km, the 1st PN corrections will be about 1000 times
smaller than the orbital corrections due to the change in
binding energy discussed in this paper.
The model of mass transfer considered in this paper
makes use of the Roche model, ignores the spin angular
momentum of the secondary, and calculates the angular
momentum loss by gravitational radiation in the point par-
ticle approximation. The validity of these assumptions has
been examined by Rezzolla et al. (2001) for the case of a
white dwarf primary. Rezzolla et al. (2001) do an accu-
rate numerical integration to compute the structure of a
semi-detached compressible co-rotating binary in hydro-
static equilibrium. They find that Paczynski's approxima-
tion for the Roche lobe radius and the expression for the
total angular momentum are correct to about 1% if the
secondary star's spin angular momentum is included. In
their computations, they also compute the rate that an-
gular momentum is lost when the secondary's structure
is taken into account and find that the point-particle ap-
proximation introduces an error of about 1%. While these
corrections are quite small, the relativistic corrections due
to the neutron star's structure are also quite small. For
this reason it would be desirable to consider more realistic
models such as those computed by Rezzolla et al. (2001).
Larger uncertainties lie in the assumed physics of the bi-
nary system. We need to restrict our attention to systems
where the donor star is a white dwarf, and the angular mo-
mentum loss is strictly through gravitational radiation. It
is also important that the donor star is not rapidly heated
so that it is possible to assume that the donor's radius
responds adiabatically to mass loss.
If any mass is lost from the binary system, the angular
momentum lost could potentially drive a period deriva-
tive correction which is larger than the relativistic effect
considered here. If the mass loss occurs near the neutron
star, (through the propeller mechanism, for example) the
effect is not large enough to be of importance for very low
mass companions. If mass is lost from a region near the
companion, then the large angular momentum loss would
make it impossible to measure the relativistic effect.
The gravitational-radiation-driven angular momentum
loss discussed in this paper is a continuous process, while
the mass transfer onto the star is highly episodic in that
matter can be stored in the accretion disk for years be-
fore being transfered onto the neutron star. Since our
treatment assumes a continuous transfer of mass, mea-
surements of the orbital period derivative would have to
be measured over the course of many years before there
could be any hope of measuring the relativistic parameter
γ.
The orbital frequencies typical for the binaries of inter-
est lie within the bandwidth of the Laser Interferometer
Space Antenna (LISA). The dimensionless gravitational-
wave amplitude h (Thorne 1998) measured by LISA can
be written as
10−23
24/3 f 1/3
P (cid:19)2/3(cid:18) 2kpc
D (cid:19)
h ≤
−6 (cid:18) M1
2M⊙(cid:19)4/3(cid:18) 2hr
(41)
where D is the distance to the binary. Using the measured
binary parameters for SAX J1808.4-3658, we find that the
gravitational-wave amplitude is h ≤ 10−23, which is well
below the "confusion noise" produced by the population
of galactic white dwarfs (Nelemans et al. 2001). The low
amplitude of the signal is mainly due to the small mass
ratio, so we expect that the binaries of the type discussed
in this paper will not be detectable by LISA, unless new
data analysis techniques are developed which can take ad-
vantage of the known positions of the binaries.
In order for the effects of a changing relativistic mass
deficit to be observable, very small changes in the binary's
orbital period must be measured. In addition, the rate of
mass transfer must be known in order to calculate the mag-
nitude of the relativistic effect. For this reason, we have
examined the scenario of mass transfer from a fully de-
generate dwarf driven by gravitational radiation, since the
mass transfer rate is related quite simply to the binary's
parameters. We have focused our attention on the ms
spin-period neutron stars since the prospects for measur-
8
ing the orbital period for these systems seems most promis-
ing. However, the orbital effect of a changing mass deficit
also holds for non-rotating stars. In fact, the effects are
much simpler in the case of a non-rotating star, since the
effects of torquing due to mass accretion are completely
negligible.
This research was supported by a grant from NSERC.
We would like to thank Saul Rappaport for his very useful
comments on the physics in this paper. We also thank
Symeon Konstantinidis for pointing out an error in a pre-
vious version of this paper.
Akmal, A., Pandharipande, V. R., & Ravenhall, D. G. 1998, Phys.
Markwardt, C. B., Juda, M., & Swank, J. H. 2003, IAU Circ., 8095,
REFERENCES
Rev. C, 58, 1804
Andersson, N., Kokkotas, K., & Stergioulas, N. 1999, ApJ, 516, 307
Arnett, W.D., & Bowers, R. L. 1977, ApJ Sup., 33, 415
Bardeen, J. M. 1973, "Rapidly Rotating Stars, Disks and Black
Holes" in Black Holes, Edited by C. DeWitt and B. S. DeWitt,
(Gordon and Breach, New York) 241 - 290
Bethe, H.A. & Johnson, M.B. 1974, Nucl. Phys. A 230, 1
Bildsten, L. & Chakrabarty, D. 2001, ApJ, 557, 292
Chakrabarty, D. & Morgan, E. H. 1998, Nature, 394, 346
Cook, G. B., Shapiro, S. L., & Teukolksy, S. A. 1994, ApJ, 424, 823
Deloye, C.J., & Bildsten, L. 2003, ApJ, 598, 1217
Ergma, E. & Antipova, J. 1999, A&A, 343, L45
Faulkner, J. 1971, ApJ, 170, L99
Friedman, J. L., Ipser, J. R. & Parker, L. 1986, ApJ, 304, 115
Galloway, D. K., Chakrabarty, D., Morgan, E. H., & Remillard, R.
A. 2002, ApJ, 576, L137
Ghosh, P. & Lamb, F. K. 1991, "Plasma Physics of Accreting
Neutron Stars" in Neutron Stars: Theory and Observation, Edited
by J. Ventura and D. Pines, (Kluwer, Netherlands), 363 - 444
Glendenning, G. K. 2000, Compact Stars, (Springer-Verlag, New
York)
Komatsu, H., Eriguchi, Y., & Hachisu, I. 1989, MNRAS, 237, 355
Kraft, R. P., Mathews, J., & Greenstein, J. L. 1962, ApJ, 136, 312
Nelemans, G., Yungelson, L. R., Portegies Zwart, S. F. 2001, A&A,
375, 890
Markwardt, C. B., Swank, J. H., Strohmayer, T. E., Zand, J. J. M.
in't., & Marshall, F. E. 2002, ApJ, 575, L21
2
Misner, C. W., Thorne, K. S., & Wheeler, J. A. 1973, Gravitation,
(Freeman and Co., San Francisco)
Nelson, L.A., & Rapparport, S. 2003, ApJ, 598, 431
Paczy´nski, B. 1971, ARA&A, 9, 183
Pandharipande, V.R. & Smith, R.A. 1975, Phys. Lett. B 59, 15
Podsiadlowski, P., Rappaport, S., & Pfahl, E.D. 2002, ApJ, 565, 1107
Rasio, F.A., Pfahl, E.D., Rappaport, S. 2000, ApJ, 532, L47
Rappaport, S., Fregeau, J.M., & Spruit, H. 2004, ApJ, 606, 436
Rappaport, S., Joss, P. C., & Webbink, R. F. 1982, ApJ, 254, 616
Rappaport, S., Verbunt, F., & Joss, P. C. 1983, ApJ, 275, 713
Rezzolla, L., Uryu, K., & Yoshida, S. 2001, MNRAS, 327, 888
Spyrou, N. K. 1988, J. Astrophys. Asr., 9, 25
Spyrou, N. K. & Stergioulas, N. 2001, A&A, 366, 598
Stergioulas, N. & Friedman, J. L. 1995, ApJ, 444, 306
Strohmayer, T. E., Markwardt, C. B., Swank, J. H., & in't Zand, J.
2003, ApJ, 596, L67
Thorne, K. S. 1998, "Probing Black Holes and Relativistic Stars with
Gravitational Waves" in Black Holes and Relativistic Stars Ed. by
R. Wald, (University of Chicago Press, Chicago), 41 - 78
Verbunt, F. 1993, ARA&A, 31, 93
Verbunt, F. & Zwaan, C. 1981, A&A, 100, L7
Wagoner, R. V. 1984, ApJ, 278, 345
Wang, Z., Chakrabarty, D., Roche, P., Charles, P. A., Kuulkers, E.,
Shahbaz, T., Simpson, C., Forbes, D. A., & Helsdon, S. F. 2001,
ApJ, 563, L61
Wijnands, R. & van der Klis, M. 1998, Nature, 394, 344
Observational data for accreting ms pulsar systems and values of the constants D and F .
Table 1
Binary System
νs (Hz)a
SAX J1808.4-3658
XTE J1751-305
XTE J0929-314
XTE J1807-294
XTE J1814-338
401
435
185
191
314
νs(Hz s−1)
< 7 × 10−13
< 3 × 10−13
−9.2 × 10−14
P (s)b
f (M⊙)c
7249.119
2545.3414
2614.746
2404.2
15390
3.7789 × 10−5
1.2797 × 10−6
2.7 × 10−7
Dd
2.274
4.210
1.1
Reference
F e
3.3 Chakrabarty & Morgan (1998)
5.7 Markwardt et al. (2002)
0.4 Galloway et al. (2002)
Markwardt et al (2003)
Strohmayer et al (2003)
aPulsar spin frequency
bOrbital Period
cMass function
dD is defined in equation (37).
eF is defined in equation (39).
|
astro-ph/0311312 | 1 | 0311 | 2003-11-13T14:03:03 | Accretion Driven Evolution of Quasars and Black Holes: Theoretical Models | [
"astro-ph"
] | We present a flexible framework for constructing physical models of quasar evolution that can incorporate a variety of observational constraints, such as multi-wavelength luminosity functions, estimated masses and accretion rates of active black holes, space densities of quasar hosts, and the local black hole mass function. We focus on the accretion rate distribution p(mdot|M,z), the probability that a black hole of mass M at redshift z accretes at a rate mdot in Eddington units. Given the radiative efficiency as a function of mdot, the quasar luminosity function (QLF) is determined by a convolution of p(mdot|M,z) with the black hole mass function n(M,z). In the absence of mergers p(mdot|M,z) also determines the full evolution of n(M,z), given a "boundary value" of n(M) at some redshift. Matching the observed decline of the QLF break luminosity at z<2 requires either a shift towards lower characteristic accretion rates or an evolving mass dependence of p(mdot) that preferentially shuts off accretion onto high mass black holes. These two scenarios make different predictions for the masses and accretion rates of active black holes. If the first mechanism dominates, then the QLF changes character between z=2 and z=0, shifting from a sequence of black hole mass towards a sequence of L/Ledd. We construct and compare five models that illustrate different assumptions about the quasar population: short and long lifetime models dominated by unobscured thin-disk accretion, a model with a high fraction of obscured quasars, and models in which mergers or ADAF accretion produce substantial black hole growth at low redshift. We discuss the observational advances that would be most valuable for distinguishing such models and for pinning down the physics that drives black hole and quasar evolution. (Abridged) | astro-ph | astro-ph |
Accretion Driven Evolution of Quasars and Black Holes: Theoretical Models
Adam Steed and David H. Weinberg
Department of Astronomy, The Ohio State University, Columbus, OH 43210
asteed,[email protected]
ABSTRACT
We present a flexible framework for constructing physical models of quasar evolu-
tion that can incorporate a wide variety of observational constraints, such as multi-
wavelength luminosity functions, estimated masses and accretion rates of active black
holes, space densities of quasar host galaxies, clustering measurements, and the mass
function of black holes in the local universe. The central actor in this formulation is the
accretion rate distribution p( mM, z), the probability that a black hole of mass M at
redshift z accretes at a rate m in Eddington units. Given a model of accretion physics
that specifies the radiative efficiency and SED shape as a function of
m, the quasar
luminosity function (QLF) is determined by a convolution of p( mM, z) with the black
hole mass function n(M, z). In the absence of mergers, p( mM, z) also determines the
full evolution of n(M, z), given a "boundary value" of n(M ) at some redshift. If p( mz)
is independent of mass, then the asymptotic slopes of the QLF match the asymptotic
slopes of n(M ), and n(M ) evolves in a self-similar fashion, retaining its shape while
shifting to higher masses. Matching the observed decline of the QLF "break" lumi-
nosity at z < 2 requires either a shift in p( mz) that increases the relative probability
of low accretion rates or an evolving mass dependence of p( mM, z) that preferentially
shuts off accretion onto high mass black holes at low z. These two scenarios make
different predictions for the masses and accretion rates of active black holes.
If the
first mechanism dominates, then the QLF changes character between z = 2 and z = 0,
shifting from a sequence of black hole mass towards a sequence of L/LEdd. We use our
framework to compare the predictions of five models that illustrate different assump-
tions about the quasar population: two dominated by unobscured thin-disk accretion
with short and long quasar lifetimes, respectively, one with a 4:1 ratio of obscured to
unobscured systems, one with substantial black hole merger activity at low redshift,
and one with substantial low redshift growth in radiatively inefficient flows. We discuss
the observational advances that would be most valuable for distinguishing such models
and for pinning down the physics that drives black hole and quasar evolution.
Subject headings: quasars: general
Submitted to The Astrophysical Journal, November 19, 2018.
-- 2 --
1.
Introduction
The study of the quasar and AGN populations has been transformed in recent years by am-
bitious new optical (e.g., Boyle et al. 2000; Schneider et al. 2002; Wolf et al. 2003), X-ray (e.g.,
Brandt et al. 2001; Giacconi et al. 2002; Anderson et al. 2003; Ueda et al. 2003), and radio (e.g.,
White et al. 2000) surveys, by the recognition that low efficiency accretion modes may become
important when the accretion rate itself is low (e.g., Narayan et al. 1998 and references therein),
by detailed studies of low luminosity AGN in the local universe (e.g., Ho 2001), and, perhaps most
of all, by the accumulating evidence that supermassive black holes are ubiquitous in the bulges of
present-day galaxies (e.g., Richstone et al. 1998; Merritt & Ferrarese 2000; Gebhardt et al. 2000).
The dynamical studies of nearby galaxies strengthen the long-standing hypothesis that quasars are
powered by black hole accretion (e.g., Lynden-Bell 1969; Rees 1978), and the "demography" of
the local black hole population provides a powerful constraint on models of quasar evolution and
its connection to galaxy evolution. These developments have inspired increasingly sophisticated
theoretical models that place quasar evolution in the context of hierarchical clustering models for
the formation of dark matter halos and galaxies (e.g., Kauffmann & Haehnelt 2000; Cavaliere &
Vittorini 2002; Haiman & Loeb 2001; Wyithe & Loeb 2003).
This paper presents a physically motivated calculational framework that is intermediate in
complexity between such ab initio models of the quasar population and older descriptions in terms
of "luminosity evolution" or "density evolution." The central actor in our formulation is the
accretion probability distribution p( mM, z), the probability that a black hole of mass M at redshift
z is accreting mass at a rate m in Eddington units (discussed in §2 below). The key supporting
players are the black hole mass function n(M, z) and a physical model of accretion that predicts
the radiative efficiency for a given m.1 An example of an accretion model would be thin-disk
accretion with efficiency ǫ ≡ L/ M c2 ∼ 0.1 when m ∼ 1, changing to low efficiency advection-
dominated (ADAF) accretion when m is below some critical value. At a given redshift, p( mM )
and n(M ) together determine the quasar luminosity function, and p( mM ) also determines the
accretion driven growth of the black hole population, and hence the evolution of n(M ). Thus, given
physical assumptions about radiative efficiencies and a "boundary condition" specifying n(M ) at
one redshift, the history of p( mM ) determines the complete evolution of the black hole population
and the quasar luminosity function. An essential caveat is that mergers of black holes following
mergers of their host galaxies could alter n(M ) independently of the accretion characterized by
p( mM ).
The simplest scenario connecting black hole and quasar evolution is that black holes "shine as
they grow": a luminous quasar is powered by a black hole radiating at near-Eddington luminosity
with efficiency ǫ ∼ 0.1, and no significant growth occurs in a non-luminous phase. In this case,
the bolometric luminosity function at a given redshift is just Φ(L) = fonn(L/l)l−1 , where l is the
1Henceforth, we will usually drop the explicit dependence on z and refer only to p( mM ) or n(M ), but these
should always be understood to refer to the distribution at some particular redshift.
-- 3 --
(universal) ratio of Eddington luminosity to black hole mass and fon is the fraction of black holes
that are accreting. In a time interval ∆t, black holes on average increase their mass by a factor
exp(fon∆t/tg), where tg = M/(LEdd/ǫc2) = 4.5 × 107(ǫ/0.1) yr is the e-folding time for growth at
the Eddington luminosity (Salpeter 1964; see discussion in §2). The mass density in black holes
at the present day is simply related to the emissivity of the quasar population integrated over
0 U (t)dt/ǫc2 (Soltan 1982, updated by, e.g., Chokshi & Turner
1992; Richstone et al. 1998; Yu & Tremaine 2002; Fabian 2003). In our language, this is a model
in which all active black holes have the same radiative efficiency and p( mM ) = fonδD( m − 1),
where δD is the Dirac-delta function. The evolution of quasars and black holes is determined by
a boundary condition on n(M ) and the redshift history of the active fraction fon(z). The "quasar
era" z ∼ 2 − 4 when the emissivity of the population peaks is also the era in which today's black
holes grew to their current mass.
luminosity and redshift, ρBH = R t0
This simple scenario may not be too far from the truth, but the possible complications raise a
number of questions. Did today's black holes gain a significant fraction of their mass through low-
m, low efficiency, ADAF-type accretion, thus growing at low luminosity? Have black hole mergers
substantially altered n(M ), leaving the integrated density ρBH fixed but changing the relative
numbers of high and low mass black holes? Are some quasars accreting mass at super-Eddington
rates, radiating at high luminosity but low efficiency in "smothered," optically thick ADAF modes
(Katz 1977; Begelman 1978; Abramowicz et al. 1988)? Do some quasars radiate substantially above
the Eddington limit (Begelman 2002)? Is a significant fraction of quasar activity obscured by gas
and dust, as hypothesized in synthesis models of the X-ray background (e.g., Setti & Woltjer 1989;
Comastri et al. 1995; Fiore et al. 1999; Fabian & Iwasawa 1999; Gilli et al. 2001), thus redistributing
bolometric luminosity from the optical-UV-soft X-ray to the far-IR? Are low luminosity AGNs
powered mainly by low mass black holes radiating at Eddington luminosity, by more massive black
holes with thin-disk efficiencies but sub-Eddington accretion rates, or by still more massive black
holes with sub-Eddington accretion rates and low efficiency? The methods developed here provide
useful tools for addressing these questions, allowing us to construct concrete, quantitative models
that answer them in different ways, then examine how observational data might distinguish among
such models.
Our framework complements, but by no means replaces, the ab initio approach that connects
the evolution of quasars and black holes to that of the underlying dark halo and galaxy populations.
This approach has yielded many valuable insights, including the recognition that the rise of the
quasar population probably traces the formation of the first dark halos large enough to host massive
black holes, that the rapid decline of the population at low redshift probably reflects the combined
impact of declining galaxy interaction rates and decreasing gas supplies in quasar hosts, and that the
clustering of quasars with themselves or with galaxies can provide a valuable diagnostic of typical
quasar lifetimes (e.g., Efstathiou & Rees 1988; Haehnelt & Rees 1993; Haehnelt et al. 1998; Salucci
et al. 1999; Kauffmann & Haehnelt 2000; Cavaliere & Vittorini 2000; Haiman & Hui 2001; Martini
& Weinberg 2001; Menou et al. 2001; Wyithe & Loeb 2003). In combination with semi-analytic
-- 4 --
models of galaxy formation, these quasar evolution models can also predict the properties and
environments of quasar hosts and the relation between the properties of present day galaxies and
the masses of their central black holes. However, the models necessarily rely on specific assumptions
about the mechanisms that trigger quasar activity and the accretion rates that these mechanisms
produce. To put things in our terms, the ab initio models adopt a particular set of hypotheses
about quasar activity in order to predict p( mM, z) from first principles.
Our framework is designed to model observational data in a flexible way with relatively few as-
sumptions, while retaining the basic physical picture of black hole accretion that underlies nearly all
modern interpretations of the quasar population. One hope is that measurements of the luminosity
function and the local black hole mass function will eventually allow us to integrate backwards in
time and determine p( mM, z) empirically, drawing on a variety of observations to test the assump-
tions that enter such a reconstruction. We may find that the data are not powerful enough to tie
down p( mM, z) without some a priori constraints on its expected form, but that finding in itself
would be a valuable, if disappointing, lesson. More generally, we hope to illuminate the connections
between black hole evolution and quasar activity and learn what observations can and cannot tell
us about these connections.
The Haehnelt et al. (1998) paper has had the strongest impact on our thinking about these
issues, but the most direct antecedent that we know of to the approach taken here is the lucid paper
of Small & Blandford (1992). They adopted a similar description of black hole evolution and its
connection to quasar activity, and they applied this description to the observational data available
at the time. Advances in the observational data and the theoretical models of accretion make
this an opportune time to revive and extend this approach. Yu & Tremaine (2002) have recently
used a similar method in assessing constraints on black hole accretion and mergers, though their
assumptions and goals are more restrictive than ours -- in particular, they assume that quasars
radiate at Eddington luminosity and thus that p( m) consists of δ-functions at m = 1 and m = 0.
The difference in the form of p( m) is one of the most significant differences between the
models presented in this paper and most models of the quasar population in the literature. These
typically assume that p( m) for m > 0 is sharply peaked at some value close to Eddington, such
as a δ-function (e.g., Small & Blandford 1992) or a spike followed by an exponential decline (e.g.,
Haehnelt et al. 1998; Kauffmann & Haehnelt 2000). A sharply peaked p( m) could arise physically
if the central black hole typically plays a large role in controlling its own fuel supply, through
feedback or influence on stellar dynamics. However, we think it is more likely that fueling is driven
by galactic scale events -- galaxy mergers, interactions, and bar formation, for example -- that
are minimally influenced by the central black hole, and therefore do not "know" that they should
feed it at any particular rate. In particular, it seems reasonable that for every major event that
leads to Eddington-like fueling of a central black hole there are many minor events that fuel it
at a sub-Eddington rate. The particular functional forms that we adopt here to represent this
scenario, a power-law or broken power-law p( m) between some mmin and some mmax, are arbitrary,
and chosen largely for mathematical convenience, but they reflect this general thinking about the
-- 5 --
process of quasar fueling. In the long run, one goal of our effort is to test observationally whether
p( m) is in fact a broad function or a peaked function, which would in turn have implications for
the mechanisms of quasar fueling. At a qualitative level, the wide L/LEdd distribution of AGN
activity in the local universe (e.g, Ho 2001) seems to support the idea of a broad distribution of
accretion rates.
In this paper we will keep our contact with observations relatively loose, focusing instead
on presenting our framework in a clear way and illustrating how observations might distinguish
among different scenarios for the quasar population. We will carry out a detailed analysis of multi-
wavelength measurements of the quasar luminosity function and constraints on active black hole
masses and the local black hole mass function a subsequent paper. We adopt a cosmological model
with Ωm = 1 and h ≡ H0/100 km s−1 Mpc−1 = 0.5 because all of the observational papers include
results for this model, and not always for the low density, Λ-dominated, h ∼ 0.7 model favored by
recent cosmological observations. Cosmology affects our models indirectly through its influence on
observationally inferred luminosity functions and directly through the time-redshift relation. We
would not expect a change of cosmological model to have any qualitative impact on our results, and
we expect that the quantitative impact could be compensated by modest changes to mass scales
and accretion rates, especially since the age of the universe is similar for Ωm = 1, h = 0.5, and for
Ωm = 0.3, ΩΛ = 0.7, h = 0.7.
We present the definitions and key equations of our framework in the following section. We
then present mathematical results for luminosity functions in §3 and for luminosity and black
hole mass evolution in §4, focusing on analytically solvable cases that illustrate general points.
In §5 we construct five illustrative models of the quasar population, each designed to match the
observed evolution of the optical luminosity function but differing from one another in the typical
quasar lifetime or in the importance of mergers, ADAF growth, or obscuration. We show how
measurements of the black hole mass function, luminosity functions at other wavelengths, the
masses and accretion rates of active black holes, and the space density of host galaxies might
distinguish among these scenarios. We summarize our results in §6. This is a long paper, and
a reader who wants to get quickly to the main points can read §2, look through the figures and
captions, and read §6. The definitions of the models illustrated in Figures 8 -- 17 are summarized in
Table 2 and its accompanying caption.
2. Framework
2.1. Definitions and assumptions
We define l to be the ratio of a black hole's Eddington luminosity to its mass,
l ≡
LEdd
M
=
4πGmpc
σT
= 1.26 × 1038erg sec−1M⊙
−1.
(1)
We often scale the radiative efficiencies to the value 0.1 that is typically adopted for thin-disk
accretion,
ǫ ≡
= 0.1ǫ0.1,
(2)
-- 6 --
L
M c2
where M is the mass accretion rate and L is the bolometric luminosity. We define the Eddington
accretion rate to be the mass accretion rate for which a black hole with radiative efficiency ǫ0.1 = 1
has Eddington luminosity,
MEdd ≡
and the dimensionless accretion rate
LEdd
0.1c2 ≈ 22(cid:18) M
109M⊙(cid:19) M⊙ yr−1,
m ≡
M
MEdd
.
(3)
(4)
This definition of
universal. A black hole accreting at
MEdd in terms of a fixed, thin-disk efficiency is common in the literature, but not
MEdd grows in mass exponentially, with an e-folding timescale
ts ≡
M
MEdd
= 4.5 × 107yr.
(5)
For ǫ0.1 = 1, ts is equal equal to the Salpeter (1964) timescale for growth at the Eddington lumi-
nosity; note, however, that we define ts to be 4.5 × 107 yr independent of efficiency and of L/LEdd.
With these definitions, a black hole of mass M accreting at a dimensionless rate m with radiative
efficiency ǫ = 0.1ǫ0.1 has bolometric luminosity
L = 0.1ǫ0.1
M c2 = ǫ0.1 mlM .
(6)
At each redshift, we define the black hole mass function n(M ) such that n(M )dM is the
comoving space density of black holes in the mass range M → M + dM ; the units of n(M ) are
thus number per comoving volume per unit mass.
In our plots, we usually show M n(M ), the
comoving space density (with units Mpc−3) in an interval d ln M , rather than n(M ) itself. We
define the accretion probability distribution p( mM ) such that p( mM )d m is the probability that
a black hole of mass M has an accretion rate in the range m → m + d m.
In our subsequent
calculations, we will frequently consider the restricted and analytically convenient class of models
in which p( mM ) = p( m), i.e., with accretion probability distribution independent of mass. Since
more massive black holes reside in more massive galaxies that have larger internal gas supplies and
can cannibalize larger companions, this assumption could be a reasonable approximation to the real
universe (see further discussion in §4.2.2 below). However, it is at best a convenient approximation,
and we will try to be mindful of its limitations.
Many theoretical models of quasar evolution specify p( m) implicitly through a typical "light
curve" L(t) that follows each triggering event. The probability p( m)d m can be identified with the
fraction of time that L/LEdd is in the range ǫ0.1 m → ǫ0.1( m + d m). However, the relation between
-- 7 --
light curves and p( m) distributions is many-to-one, even for constant efficiency. For example, if
every quasar lights up at the Eddington luminosity and declines exponentially thereafter, with
ǫ0.1 = 1 throughout, then there is constant accretion probability per logarithmic interval of
m, and
thus p( m) ∝ m−1 for m ≤ 1. But the same p( m) could correspond to an "on-off" activity model
where the luminosity of each quasar is constant for a fixed time while the number of quasars per
luminosity interval is proportional to (L/LEdd)−1.
The model of accretion physics specifies the probability that a black hole of mass M and ac-
cretion rate m has bolometric radiative efficiency ǫ0.1. Since most of the properties of accretion
flows scale in a fairly simple way with mass, it is reasonable to expect that this probability dis-
tribution is independent of M . To simplify our calculations, we will also assume that all black
holes of the same m have the same ǫ0.1, i.e., we will assume that the probability distribution of ǫ0.1
has zero width at given m. Outside of a small range of
m where black holes might cycle between
thin-disk accretion and an ADAF-type flow, this assumption seems plausible, though one could
imagine that a range of black hole spins could induce a range of ǫ0.1 values even at fixed m. We
will generally assume that ǫ0.1 = 1 in a range mcrit ≤ m ≤ 1, where mcrit is a critical value of the
accretion rate below which thin-disk accretion gives way to some lower efficiency mode. For stellar
mass black holes, which exhibit transitions among accretion modes,
mcrit may be as high as 0.09
(Esin et al. 1997), but for supermassive black holes the value is more uncertain and probably lower
(R. Narayan, private communication). We will adopt mcrit = 0.01. Below this threshold, we will
assume ǫ0.1 = ( m/ mcrit), a scaling motivated by ADAF models (Narayan et al. 1998).
If the accretion rate is determined by large scale gas flows beyond the influence of the black hole
itself, then there is no particular reason that m should not exceed one. In such situations, we will
assume that the accretion proceeds in a lower efficiency, "smothered" mode (Katz 1977; Begelman
1978; Abramowicz et al. 1988), so that the black hole radiates at the Eddington luminosity -- in
other words, ǫ0.1 = m−1 when the accretion rate is super-Eddington. It is not clear that these
smothered accretion modes are stable enough to exist in nature, and it may be that black holes
in this situation drive the excess gas out of the nucleus altogether rather than accepting mass
at a super-Eddington rate, thus regulating the accretion rate so that m never exceeds one. The
distinction between these two pictures is usually not important for our purposes, provided that
the black holes radiate at the Eddington luminosity in either case, but it does make a difference if
super-Eddington accretion makes a significant contribution to black hole mass growth. It has also
been suggested (Begelman 2002) that black holes can radiate substantially above the Eddington
luminosity (an order of magnitude or more) -- this would make a difference to our predictions, as
we discuss briefly in §6.
To summarize, we generally assume that the bolometric radiative efficiency is
( m/ mcrit)
for m < mcrit
1
m−1
for mcrit ≤ m ≤1
for m > 1,
ǫ0.1( m) =
(7)
-- 8 --
with mcrit = 0.01. We will also usually assume that p( m) is non-zero only over some range mmin
to mmax, and in some simplified cases we will choose this range so that only the thin-disk regime
with ǫ0.1( m) = 1 contributes. To calculate luminosities in particular wavebands, we will incorporate
luminosity fractions Fν , which in some cases we allow to depend on m or to vary from one black
hole to another. We discuss our assumptions about Fν as they arise.
2.2. The luminosity function and the black hole evolution equation
The QLF is obtained from the black hole mass function and the accretion probability distri-
bution by a straightforward counting argument. Black holes in the mass range M → M + dM with
accretion rate m correspond to quasars with luminosity in the range L → L + dL with L = ǫ0.1 mlM
and dL = (ǫ0.1 ml)dM (eq. 6). The comoving space density of black holes in this mass range is
n(M )dM , and the corresponding density of quasars with luminosity L → L + dL, is obtained by
multiplying by the accretion probability p( mM ) and integrating over m:
Φ(L)dL =Z mmax
mmin
d m p( mM )n(M )
dL
ǫ0.1 ml
,
M =
L
ǫ0.1 ml
,
(8)
where the integral covers the full range over which p( mM ) is non-zero. One can obtain a mathe-
matically equivalent expression by identifying the luminosity range dL with the range of accretion
rates d m = dL/(ǫ0.1lM ) at fixed M , then integrating over masses. The above argument is slightly
complicated by the possibility that ǫ0.1 depends on m, but provided that ǫ0.1 m is a monotonic
function of
m, the probability density transformation p( m)d m = p(ǫ0.1 m)d(ǫ0.1 m) still leads to
equation (8). For our standard assumption about super-Eddington accretion, ǫ0.1 m = 1 for m > 1,
the luminosity function of super-Eddington accretors is
ΦSE(L)dL = n(L/l)
dL
l Z mmax
1
d m p( mM = L/l),
(9)
so equation (8) remains valid when mmax > 1. Allowing a range of efficiencies at fixed m would
broaden the luminosity function, since one would then convolve n(M ) with p(ǫ0.1 m) instead of p( m)
to obtain Φ(L).
Equation (8) gives the bolometric luminosity function in terms of n(M ) and p( mM ) at a
If accretion is the only process contributing to black hole growth, then the
particular redshift.
evolution of n(M ) is determined by a simple continuity equation,
∂n(M, t)
∂t
= −
∂(nh M (t)i)
∂M
= −
1
ts
∂(nM h m(M, t)i)
∂M
,
(10)
which follows from considering the rate at which black holes are leaving and entering a mass range
M → M + dM (Small & Blandford 1992). (The last equality follows from equations [4] and [5],
M = mM/ts.) The important simplification that follows from the continuity
which imply that
-- 9 --
argument is that the evolution of n(M ) depends on p( mM ) only through the mean accretion rate,
h m(M, t)i =R d m m p( mM, t). In any given time interval, some black holes will grow faster than
average and some will grow slower, but the mass function is an average over the full population,
and its evolution depends only on the average rate at which black holes move from one mass range
to another.
If h m(t)i is independent of M , then it can be moved out of the derivative on the r.h.s. of
equation (10), and in this case the solution to the evolution equation is remarkably simple:
n(M, t) = F (cid:18) M
M∗(cid:19) M∗,i
M∗
,
M∗(t) = M∗,i exp(cid:18)Z t
ti
h m(t)i
dt
ts(cid:19) ,
(11)
where F (x) is an arbitrary function and M∗ is any fiducial scale in the mass function. This solution
can be verified by direct substitution into equation (10), but its physical basis is easy to see. Since
h m(t)i is independent of mass, and only h mi matters rather than the full distribution p( m), the
evolution is equivalent to that of a population in which all black holes accrete mass at the rate
M = h m(t)i MEdd = h m(t)iM/ts. In this case, every black hole grows by the exponential factor on
the r.h.s. of equation (11), so the scale of the mass function simply shifts; the normalization drops
in proportion to 1/M∗ because the width of the differential mass range dM occupied by a given
set of black holes increases as the black hole masses themselves increase. The simplicity of the
solution (11) makes models in which p( m) is independent of M more tractable than others, so this
restricted class of models is useful for gaining intuition and illustrating different types of behavior.
The more general equation for the evolution of n(M, t) is
∂n(M, t)
∂t
= −
1
ts
∂(nM h m(M, t)i)
∂M
+ C(M, t) − D(M, t)
dM ′n(M − M ′, t)n(M ′, t)Γ(M − M ′, M ′, t)
0
+ Z M
− n(M, t)Z ∞
dM ′n(M ′, t)Γ(M, M ′, t) ,
(12)
0
where the last two terms represent formation of black holes of mass M by mergers of black holes
of mass M ′ and M − M ′ and loss of black holes of mass M by mergers with other black holes, and
the creation (C) and destruction (D) terms allow for processes that are neither smooth accretion
nor mergers. Equation (12) is essentially the same equation that Murali et al. (2002) use to model
the evolution of the galaxy mass function by accretion and mergers (see their equation A1). The
genuine destruction of supermassive black holes seems an unlikely prospect, but they could be
removed from the population of galactic nuclei by ejection in three-body encounters (e.g., Valtonen
et al. 1994), which would effectively count as destruction for our purposes. We will not consider
this possibility in our models here, but it could be a significant effect if there is a long delay between
the merger of galaxies and the merger of their central black holes (see, e.g., Madau et al. 2003).
We will also assume that there is no black hole creation in the mass and redshift range that we
consider, i.e., all of the black holes are already present at the highest redshift in our calculations,
-- 10 --
and their masses change only by accretion and mergers. A calculation that included the formation
of "seed" black holes would need to incorporate the creation term C(M, t), but here we treat the
mass distribution of these seeds as the initial condition for evolution of n(M ). In the context of
our models, a "seed" black hole is one that forms at low efficiency by a process that is not well
described by the same p( m) governing most accretion driven growth -- e.g., by the collapse of a
supermassive star, a relativistic gas disk, or a relativistic star cluster.
If the accretion, creation, and destruction terms are ignored, equation (12) becomes the "co-
agulation equation," which has been widely studied in the context of planetesimal growth (see Lee
2000 and references therein) and applied on occasion to star formation and galaxy clustering (e.g.,
Silk & White 1978; Silk & Takahashi 1979; Murray & Lin 1996; Sheth 1998). Unfortunately, the
coagulation equation has analytic solutions only for rather specialized forms of the collision rates
Γ(M, M ′, t) that do not seem particularly applicable to black hole evolution, and even these solu-
tions no longer apply once accretion is also important. The usefulness of equation (12) is therefore
largely conceptual, and as a basis for numerical calculations given some physically motivated forms
of the collision rates. In this paper, we will restrict our consideration of mergers to idealized recipes
that are, we hope, sufficient to illustrate their generic effects.
3. Luminosity Functions
3.1. Power-Law p( m) for m in the Thin-Disk Range
The functions p( mM ) and n(M ) could in principle have complicated forms, but it proves
useful to consider some simple forms and determine qualitative behaviors that would hold true in
more general cases. We begin by implementing our framework in a specific case where we consider
only the range of accretion rates that corresponds to thin-disk accretion, i.e.
mcrit < m < 1
and thus ǫ0.1 = 1. We also assume that the probability that a black hole accretes at a rate m is
independent of its mass, i.e. p( mM ) = p( m). We first consider a truncated power-law,
p( m) = p∗ ma,
mmin < m < mmax.
(13)
[There is also a δ−function at m = 0, representing inactive black holes, so that p( m) integrates to
one.] In this section, we adopt mmin = mcrit = 0.01 (see §2.1) and mmax = 1. We adopt a broken
power-law form for the black hole mass function,
n(M ) =
M∗(cid:17)α
n∗(cid:16) M
M∗(cid:17)β
n∗(cid:16) M
M < M∗,
M > M∗.
(14)
Henceforth, we will frequently refer to the normalization of n(M ) in terms of n∗M∗, the number
density of objects within a logarithmic interval around M∗.
The convolution integral (8) for the luminosity function breaks into three regimes because the
range of accretion rates is finite. Luminosities L < ǫ0.1 mminlM∗ = 0.01lM∗ cannot be generated by
-- 11 --
black holes with masses M > M∗ because they would require an accretion rate below mmin. Thus,
only the low mass end of the black contributes to this luminosity regime,
Φ(L) =Z mmax
mmin
p∗ man∗(cid:18)
L
ǫ0.1 mlM∗(cid:19)α
1
ǫ0.1l m
d m,
L < 0.01lM∗.
(15)
Similarly, luminosities L > ǫ0.1 mmaxlM∗ = lM∗ cannot be generated by black holes with masses
M < M∗, so the integral for this regime is the same as equation (15) but with α replaced by β. For
the intermediate regime, black holes with masses above and below M∗ contribute. It is convenient
to define the accretion rate mL at which an M∗ black hole has luminosity L,
mL ≡
L
ǫ0.1lM∗
,
(16)
so that the Φ(L) integral breaks into pieces contributed by the high and low ends of the black hole
mass function:
L
Φ(L) = Z mL
+ Z mmax
mmin
mL
p∗ man∗(cid:18)
p∗ man∗(cid:18)
ǫ0.1 mlM∗(cid:19)β
ǫ0.1l mM∗(cid:19)α
L
1
ǫ0.1l m
d m
1
ǫ0.1l m
d m,
0.01lM∗ < L < lM∗.
(17)
The solution for the QLF is
Φ(L) =
n∗p∗
l (cid:20)
β−α
lM∗(cid:17)a
(a−β)(a−α) (cid:16) L
1−0.01a−α
n∗p∗
l
− 0.01a−β
lM∗(cid:17)α
(a−α) (cid:16) L
lM∗(cid:17)β
a−β (cid:16) L
lM∗(cid:17)β
(a−β) (cid:16) L
n∗p∗
1−0.01a−β
l
+ 1
a−α(cid:16) L
lM∗(cid:17)α(cid:21)
L < 0.01lM∗,
0.01lM∗ < L < lM∗,
L > lM∗,
(18)
where the values mmin = 0.01, ǫ0.1 = 1, and mmax = 1 have been explicitly included. Note that we
have scaled luminosities to the Eddington luminosity lM∗ of an M∗ black hole.
The most comprehensive observational analysis of the optical quasar luminosity function at
z . 3 is that of Boyle et al. (2000), based on the 2dF Quasar Redshift Survey and the Large Bright
Quasar Survey. They find that the rest-frame B-band luminosity function can be adequately fit by
a double power-law function of the form
Φ(LB) ∝"(cid:18) LB
Lbrk(cid:19)αL
+(cid:18) LB
Lbrk(cid:19)βL#−1
,
(19)
with the break luminosity Lbrk evolving with redshift. (We denote the break luminosity Lbrk rather
than L∗ to keep clear the distinction between this observational quantity and the characteristic
parameters of our models, M∗ and m∗.)
-- 12 --
Fig. 1. -- QLF for models with a double power-law black hole mass function and a single power-law
p( m), with thin-disk accretion only. Left panels show the input p( m) (dotted line, top and right
axis labels, dimensionless) and mass function (solid line, bottom and left axis labels). In all Figures,
we plot M n(M ), the number per comoving Mpc3 per ln M interval, rather than the mass function
n(M ) itself. Right panels show the corresponding B-band luminosity function, plotted in number
per comoving Mpc3 per magnitude against B-band absolute magnitude. Solid lines show the total
luminosity function, while long-dashed, short-dashed, and dotted lines show the contributions from
accretion rates in the ranges 0.01 < m < 0.1, 0.1 < m < 0.25, and 0.25 < m < 1.0, respectively.
Open circles show the double power-law fit to the Boyle et al. (2000) QLF measurements at z = 2,
over the range of absolute magnitudes probed by the data. Top and bottom rows show results for
p( m) power-law slopes a = 0 and a = −1.2, respectively, for the same black hole mass function.
This and all subsequent figures assume an Ωm = 1 cosmology with h = 0.5.
-- 13 --
Figure 1 shows QLFs computed by equation (18) for two different choices of the p( m) index,
a = −1 (top) and a = −2 (bottom). Left hand panels show the input mass functions and p( m), and
right hand panels show the corresponding QLF. Instead of the mass function n(M ) itself, we plot
M n(M ), the space density of black holes per ln M interval, since this quantity is easier to interpret.
We adopt n(M ) slopes α = −1.5 and β = −3.4 to match the low and high end slopes of the Boyle
et al. (2000) QLF. Note that the slopes of the QLFs in Figure 1 are actually α + 1 and β + 1 because
they are plotted in terms of magnitudes rather than luminosities. The QLFs have been converted
from bolometric emission into absolute B-band magnitudes by using Lbol/νBLB = 10.4 (Elvis et
al. 1994) to calculate the B-band luminosity and then using MB = −2.5(log LB − 32.67) + 5.48 to
convert the luminosity into an absolute magnitude, where 5.48 is the absolute magnitude of the sun
in the B band and 1032.67 erg s−1 is the solar luminosity in the B band. However, it is important to
keep in mind that these are scaled bolometric luminosity functions and correspond to true optical
luminosity functions only if this Lbol/νBLB ratio is constant from quasar to quasar (see §3.5). The
points in the right hand panels correspond to Boyle et al.'s fit (eq. 19) at z ∼ 2, and they cover
only the range of magnitudes actually observed at this redshift.
mmin
that this probability, R mmax
The amplitude of the QLF is directly proportional to p∗ and to the mass function normalization
n∗. The value of p∗ determines the probability that a given black hole will be "on" at some non-
zero luminosity. For the a = −2 model (lower panel), we have somewhat arbitrarily chosen p∗ so
p( m)d m, is equal to one. Though all of the black holes are active in
this case, most of the activity is at accretion rates near mmin. We then choose a mass function
normalization n∗M∗ = 4.6 × 10−5Mpc−3 so that the model QLF matches the Boyle et al. (2000)
data point at the break luminosity. For the a = −1 model, we have kept the same n(M ) and again
chosen p∗ to match the observed QLF at the break. Note that equation (18) reveals a complete
degeneracy between n∗ and p∗ in the QLF. Thus, at a fixed time, it is impossible to determine
from the QLF alone if there is a high number density of black holes of which a small fraction are
accreting or a low number density of black holes of which a large fraction are accreting. We will
see later that the evolution of the QLF breaks this degeneracy to some degree.
The QLF in the upper right panel of Figure 1 demonstrates one of the important general
features of this solution: the slopes of the low and high luminosity ends of the QLF are determined
by the slopes of the low and high mass ends of the black hole mass function (see eq. 18). This
behavior follows whenever p( m) is independent of M and the range of accretion rates is finite.
This result can be understood by considering that a small range d m at
m = 1 gives Φ(L) =
n(M )(ǫ0.1l)−1p( m = 1)d m with M = L/(ǫ0.1l), which is a simple mapping of the black hole mass
function. The same range d m at a lower m gives a contribution to the QLF mapped to luminosities
fainter by a factor of
m and up or down by a factor proportional to p( m)/p( m = 1). The total
QLF is just the sum of these transformed black hole mass functions. Since we have low and high
luminosity regimes in which only one slope of n(M ) contributes, the sum in these regimes will be
a sum of power-laws with the same slope, resulting in a QLF with the same slope. The mid-range
luminosity is more complicated because both parts of the black hole mass function are contributing.
-- 14 --
The relative contributions from the low and high end of the black hole mass function depend on
the relative probability of low and high accretion rates. This can be seen in the middle part of
equation (18), which shows that the slope of the probability distribution as well as the slopes of
n(M ) affect the shape of the QLF in this luminosity regime. The agreement of asymptotic slopes
between n(M ) and Φ(L) motivates our choice of a double power-law n(M ), though we will see in
§5.2 that this choice has some problems when compared to local observations.
The a = −1 model in Figure 1 corresponds to equal probabilities of accretion in each logarith-
m between mmin and mmax, but the emissivity of the population is dominated by
mic interval of
accretion rates close to mmax because luminosities are proportional to m. The model QLF is in rea-
sonably good agreement with the Boyle et al. (2000) data. Dotted, short-dashed, and long-dashed
lines in the QLF panels show the contribution from accretion rates in the ranges 0.01 < m < 0.1,
0.1 < m < 0.25, and 0.25 < m < 1.0 respectively. For a = −1, the QLF is dominated by black
holes with near-Eddington accretion rates at all luminosities. The curves for lower accretion rates
are shifted horizontally to lower luminosities, with slight vertical shifts because the three bins are
not equal logarithmic intervals.
The a = −2 model has equal contributions to the emissivity from each logarithmic interval of
m. The slow transition between the low and high luminosity regimes yields a worse fit to the Boyle
et al. (2000) QLF. High accretion rates still dominate the high end of the QLF, but low accretion
rates dominate the low end.
In general, low m black holes can more easily make a significant
contribution below the break in the luminosity function because a shift "left" can be more easily
compensated by a shift "up," especially when the slope of the low end of the mass function is
shallow. However, above the break in luminosity, it is difficult for low m accretors to make a
contribution. For example, a p( m) slope a ∼ −3.3 would be required to make the contribution of
low and high accretion rates comparable at high luminosities in Figure 1.
3.2. Double Power-Law p( m) with ADAF and Super-Eddington Accretion
Our assumptions for this section are similar to those of §3.1, except that we consider a wider
range of allowed accretion rates and adopt a double power-law form of p( m). Specifically, we
consider accretion rates in the range mmin = 10−4 < m < mmax = 10, which allows for ADAF, thin-
disk, and super-Eddington accretion modes, for which we adopt the efficiencies given in equation (7).
The functional form of n(M ) we consider here is the double power-law expressed in equation (14),
and the form of p( m) is analogous,
m < m∗,
m > m∗,
(20)
where m∗ is the characteristic accretion rate at the break in p( m). The solution for the QLF in
this case is given in Appendix A.
p( m) =
m∗(cid:17)a
p∗(cid:16) m
m∗(cid:17)b
p∗(cid:16) m
-- 15 --
Fig. 2. -- Like Fig. 1, but for double power-law p( m) covering a wider range of accretion rates,
including values in the ADAF regime and super-Eddington regime. Efficiencies as a function of
m
are calculated according to eq. (7). In the right panels, solid lines show the total QLF, and other
lines show the contributions from the ranges m > 1 (super-Eddington, long-dashed), 0.3 < m < 1
(thin-disk, high accretion rate, short-dashed), 0.01 < m < 0.3 (thin-disk, low accretion rate,
dotted), and 10−4 < m < 0.01 (ADAF, dot-dashed). Relative to the case in the top row, the middle
and bottom rows show cases with a higher characteristic accretion rate ( m∗ = 1 vs.
m∗ = 0.5) and
a steeper slope at low accretion rates (a = −1.1 vs. a = −0.5), respectively.
-- 16 --
Figure 2 is analogous to Figure 1, but for the double power-law p( m). In the upper panels,
we adopt a = −0.5, b = −3, and m∗ = 0.5. At high luminosity, the QLF is dominated by the
higher accretion rates in the thin-disk mode, but there is also a significant contribution from the
super-Eddington mode. At low luminosity, the contribution from higher accretion rates in the thin-
disk mode dominates over lower accretion rates in the thin-disk mode, with the super-Eddington
mode becoming much less significant. The contribution of the ADAF mode to the QLF is barely
noticeable, and this is typically the case in our models because low m and low ǫ0.1 combine to
push black hole luminosities down to ∼ 10−2 − 10−6 of Eddington, which is usually below observed
luminosities. ADAF contributions can be more significant at low redshifts and X-ray wavelengths,
as discussed in §3.5 below.
The middle panel shows the effect of increasing m∗ to 1.0, with p∗ decreased to keep Φ(Lbrk)
fixed. Super-Eddington accretion is now common enough that it dominates the high end of the
QLF, with a comparable but smaller contribution from thin-disk accretion with 0.3 < m < 1.0.
Lower luminosities are dominated by thin-disk accretion, with high accretion onto lower mass black
holes more important than low accretion onto high mass black holes.
The bottom panels show a case with m∗ = 0.5 and a low end slope of a = −1.1, which weights
p( m) more strongly to low accretion rates. The high luminosity end of the QLF in the bottom panel
has contributions similar to those in the upper panel, but the low luminosity end is now dominated
by lower accretion rates in the thin-disk mode. The turnover of the QLF in the lower panel is
less sharp than in the upper panel because of the change of the relative fraction of high and low
accretion rates contributing to the QLF over the transitional range of luminosity. The contribution
from the ADAF mode is higher than in the top panel, but it remains small at all luminosities.
As noted earlier, with the luminosity function alone there is a complete degeneracy between the
normalizations n∗ and p∗, subject only to the limitation that the fraction of black holes accreting at
a given time not exceed 100%. There is also a partial degeneracy between the characteristic values
M∗ and m∗: increasing M∗ shifts the break in the QLF to higher luminosity, but reducing m∗ makes
lower accretion rates dominate the QLF, shifting the break back down. With our assumption that
the efficiency ǫ0.1( m) decreases for m > 1, this tradeoff is limited to a factor of a few, since changes
in m∗ also affect the shape of the turnover in the QLF. We will see later that this tradeoff is even
more restricted when the evolution of the QLF is considered.
3.3. Mass Distribution of Active Black Holes
The masses of active black holes can be estimated using reverberation mapping (e.g., Wandel
et al. 1999; Onken & Peterson 2002), the widths of lines such as Hβ and CIV (e.g., Laor 1998;
Vestergaard 2004; Corbett et al. 2003), or indirectly from the properties of host galaxies (e.g.,
Dunlop et al. 2003), the variability power spectrum (Czerny et al. 2001), or the spectral energy
distribution (Kuraszkiewicz et al. 2000). The distribution of active black hole masses at a given
-- 17 --
luminosity depends on both the underlying black hole mass function n(M ) and the distribution of
accretion rates p( m). While the necessary measurements are challenging, especially at high redshift,
they can play a critical role in discriminating among models that make very similar predictions for
the luminosity function.
If there is a maximum value of the product ǫ0.1 m (e.g., unity if luminosities cannot exceed
Eddington, as in eq. 7), then black holes with M < (L/l)[(ǫ0.1 m)max]−1 cannot contribute to the
QLF at luminosity L because they cannot shine brightly enough. Conversely, black holes with
M > (L/l)[(ǫ0.1 m)min]−1 are always brighter than L, when they are active at all. The contribution
to the QLF from black holes with masses in the allowed range is the product of the number density of
black holes with mass M and the probability of having an accretion rate in the range m → m + ∆ m
that yields luminosity L → L + ∆L. Thus, for a quasar of luminosity L, the relative probability
that its black hole has mass M1 or M2, if both masses are in the allowed range, is
p(M1L)
p(M2L)
=
n(M1)p(cid:16) m = L
n(M2)p(cid:16) m = L
ǫ0.1lM1(cid:17) ∆L/M1
ǫ0.1lM2(cid:17) ∆L/M2
ǫ0.1l
ǫ0.1l
= (cid:18)M1
M2(cid:19)α−(a+1)
,
(21)
where (∆L/ǫ0.1lM ) = ∆ m. The rightmost equality applies for constant ǫ0.1 and power-law forms
of p( m) and n(M ), with slopes of a and α respectively. For this case we see that if α < a + 1 then
the rising low end of the mass function results in low mass black holes with high accretion rates
dominating the active population at luminosity L. Conversely, if α > a+1, then low accretion rates
are common enough that high mass black holes with low m dominate. A single power-law can only
approximate n(M ) or p( m) over a finite range, but this example gives insight into the more general
case and allows one to judge whether quasars of luminosity L are likely to be dominated by masses
near a break in n(M ), or by accretion rates near a break in p( m) or ǫ0.1( m). For example, the high
luminosity regime Figure 2 is dominated by Eddington luminosity black holes (near a break in ǫ0.1)
because of the steep slope of n(M ) at high masses.
For a statistical quantity that is easier to measure, it is usually desirable to integrate equa-
tion (21) to obtain the distribution of black hole masses for a specified range in luminosities. We
will show in §5 that such statistics can discriminate between models that yield similar QLFs over
the range of observed luminosities but have significantly different parameter values.
3.4. Mass Dependence of p( m)
A general mass-dependent distribution of accretion rates can be written in the form p( mM ) =
p0( m)D(M m), where p0( m) = p( mM0) at an arbitrarily chosen mass scale M0 and the function D
encodes the mass dependence, with D(M0 m) ≡ 1. To understand the potential influence of mass
dependence, we will consider the restricted case in which the function D(M ) is independent of
m,
and p( mM ) is thus a separable function. In this class of models, the relative probabilities of high
and low accretion rates are independent of mass, but the overall duty cycle can have an arbitrary
-- 18 --
mass dependence. Recall that the general expressions for the QLF (eq. 8) and the distribution
p(M L) of black hole masses at a given luminosity (eq. 21) involve p( mM ) and n(M ) only through
the product p( mM )n(M ). Therefore, for any QLF and p(M L) generated by a black hole mass
function n(M ) and a mass-independent p( m), there is a family of models with mass function
n′(M ) = n(M )/D(M ) and mass-dependent accretion rate distribution p′( mM ) = p( m)D(M ) that
predicts the same QLF and p(M L), for any choice of the function D(M ). The mass dependence
of p( m) therefore introduces a rather serious degeneracy into models of the luminosity function,
which cannot be broken by measurements of the distribution of active black hole masses.
The key difference within this class of degenerate models is the relation between the luminosity
function and the underlying black hole mass function n(M ). In particular, we have shown in §3.1
and §3.2 that for a mass-independent p( m) the low and high luminosity slopes of the QLF match
the low and high mass slopes of n(M ). This is no longer the case if p( m) depends on M . For
example, with D(M ) ∝ M x and a double power-law n(M ), the QLF has asymptotic slopes α + x
and β + x, rather than α and β. Thus, a measurement of the mass function of all black holes, not
just the active ones, is crucial to diagnosing the mass dependence of accretion rates.
Our discussion above focuses on the QLF at a particular redshift, and the degeneracy applies
if p( mM ) and n(M ) can be chosen at will. If one considers a range of redshifts over which black
holes grow by a substantial factor, then the predictions of models with different mass dependence
of p( mM ) are likely to diverge, since the mass-dependence of growth rates will change the shape
of n(M, z) from one model to the next (see §4.2.2). Thus, this degeneracy should be less serious in
a complete evolutionary model of the population. Furthermore, a measurement of n(M ) at z = 0
may be sufficient to diagnose the mass dependence of p( mM ) at higher redshift.
3.5. Wavelength Dependent Efficiencies
Equation (8) gives the bolometric luminosity function Φ(L) in terms of the accretion rate
distribution, the black hole mass function, and the efficiency ǫ0.1, which may itself be a function
of
m. If all quasars have the same spectral energy distribution (SED), then the translation to the
luminosity function in a band at frequency ν is straightforward:
Φ(Lν) =Z mmax
mmin
p( m)n(cid:18)M =
Lν
ǫ0.1 mlFν(cid:19)
1
ǫ0.1 mlFν
d m ,
(22)
where Fν ≡ Lν/Lbol is the fraction of the quasar's bolometric luminosity that emerges in the ν-
band. (Note that we are using subscript-ν to represent a finite band, not a monochromatic flux
density.) We have so far presented results for the rest-frame B-band luminosity function, assuming
that all accreting black holes have the broad-band SED estimated by Elvis et al. (1994). For a
universal SED, the luminosity functions in all bands are just shifted versions of the bolometric
luminosity function, so they all have the same shape.
The story is more interesting if some accreting black holes have radically different SED shapes.
-- 19 --
Here we will consider two representative examples, an "obscured" accretion mode in which optical,
UV, and soft X-ray radiation are absorbed by gas and dust near the nucleus and re-radiated in
the far-IR, and an ADAF mode that has a high ratio of X-ray flux to optical flux (in addition
to a reduced bolometric efficiency). Obscured accretion is thought to play an important role in
producing the X-ray background, and typical synthesis models in the literature have a ∼ 4 : 1 ratio
of obscured to unobscured sources (e.g., Comastri et al. 1995; Fabian & Iwasawa 1999). More
recent results from Chandra show that the obscured fraction is probably lower than this, especially
at high luminosities (e.g., Barger et al. 2002; Ueda et al. 2003). ADAFs are expected on theoretical
grounds to have depressed UV/optical emission relative to X-ray and far-IR (Narayan et al. 1998),
and many low luminosity AGN in the nearby universe, including Sgr A∗ in the Galactic Center,
appear to have these broader SED shapes (e.g., Ho 1999). There are, of course, other possibilities
for SED variations, including a steady change of SED shape with black hole mass caused by the
lower characteristic temperatures around higher mass black holes, a change of SED shape in the
super-Eddington regime, and perhaps a transition within the ∼ 0.1 − 1LEdd regime as the accretion
disk grows in importance relative to the hard X-ray corona.
The first task is to calculate values of Fν for the model SEDs. We will consider luminosity
functions in the rest-frame B-band, soft (0.5-2 keV) and hard (2-10 keV) X-ray bands, and a
"far-IR" band that we take to cover the range 10 − 1000µm. Since most X-ray studies work with
observed-frame fluxes (though see Cowie et al. [2003], Steffen et al. [2003], and Ueda et al. [2003]
for recent efforts to measure evolution of the rest-frame 2-8 keV luminosity function), we also
consider the soft and hard X-ray bandpasses redshifted to z = 0.5, 1, and 2. We assume that
our far-IR band is wide enough that all high-energy radiation absorbed in obscured systems is re-
radiated somewhere within it. Unfortunately, realistic experiments are likely to probe a narrower
band, for which the predictions may be quite sensitive to assumptions about dust temperatures
and departures from a blackbody spectrum. We will not consider radio luminosities here. To the
extent that the radio-quiet/radio-loud dichotomy is a simple effect of orientation or black hole spin,
it could be treated as a stochastic variation analogous to our treatment of obscuration. However,
radio luminosity may also be connected to accretion rate or to black hole mass (Dunlop et al. 2003).
Table 1 summarizes our values of Fν . We assume that unobscured quasars with accretion rates
m > mcrit = 0.01 have the mean radio quiet SED of Elvis et al. (1994), and we obtain Fν by
integrating over the appropriate wavelength bands. Simple power-law extrapolations were used in
regions without observations, and the high energy SED was extended to 30 keV by using a power-
law with photon index ΓX = 1.9. For obscured quasars, we assume the same "underlying" SED
and compute absorption in the X-rays by taking a mean X-ray photon index of ΓX = 1.89 and an
obscuring column density of NH = 3 × 1023 cm−2, which represents the median value used in the
X-ray background sythesis models of Comastri et al. (2001). We further assume that the high gas
column is accompanied by enough dust to completely extinguish the B-band flux; this assumption
may be inaccurate, as some observed systems appear to have significant optical/UV flux despite
strong absorption in the X-ray. Finally, we assume that all of the energy absorbed from the optical
-- 20 --
to the soft X-rays, about 52% of the bolometric energy in the Elvis et al. (1994) SED, is reradiated
in the FIR. This assumption seems physically plausible, though it is also possible that some or
most of the absorbed energy is channeled into driving an outflow and never radiated at all.
As in our previous calculations, we assume that accretion rates m < mcrit = 0.01 lead to
reduced efficiency, ADAF flows, but now we assign these flows a different SED. The appropriate
SED for ADAF systems is quite uncertain, and we have elected simply to take the nucleus of
M81 as representative of all black holes accreting in this mode. We calculate Fν values from the
observations tabulated in Ho (1999, tables 2 and 9 and §2.1), using the estimated Lbol and directly
observed Lν when available and otherwise using the observed monochromatic νLν values and an
appropriate αν to integrate over the waveband of interest. Since the observed FIR flux value of
M81 is an upper limit, we used the model of M81 presented in Quataert et al. (1999, fig. 1) to
estimate FFIR for the ADAF mode.
With the results of Table 1 in hand, we can again calculate multi-wavelength luminosity func-
tions using equation (22), but now the results must be computed separately for the three modes
(thin-disk, obscured, ADAF) and added together. We assume that super-Eddington accretion has
a thin-disk SED; this seems unlikely, but we do not have much idea of what to assume in its stead,
and it makes little difference to our results.
Figure 3 illustrates the potential influence of a large obscured quasar population on the multi-
wavelength QLF at z = 2. Solid lines show results for a model with no obscured quasars, with
the same black hole mass function and p( m) used in upper panels of Figure 2. Dashed lines show
a model in which these unobscured quasars are only 20% of the total -- i.e., we multiply p( m)
by four but assign 80% of the systems with m > 0.01 the obscured SED of Table 1. Since the
obscured SED has no B-band flux, the B-band luminosity function is unchanged (top left panel).
However, obscuration in the observed-frame 2-10 keV band (rest-frame 6-30 keV) is minimal, so the
hard X-ray luminosity function is nearly a factor of five higher in amplitude, with the contribution
Table 1. Fractional Bolometric Output
Thin-Disk
Obscured
ADAF
F2−10 F3−15 F4−20 F6−30
0.026
0.031
0.027
0.008
0.084
0.099
0.030
0.020
0.093
0.028
0.015
0.089
F0.5−2
0.020
1×10−9
0.057
F0.75−3
0.020
F1−4
0.020
F1.5−6
0.022
1.3×10−5
2.9×10−4
2.6×10−3
0.061
0.064
0.068
FB−band FFIR
0.17
0.79
0.03
0.025
0.00
0.012
Note. -- Values of the inverse bolometric correction, Fν−band, where Lν−band = Fν−bandLBol. Values include
intrinsic X-ray ranges at z = 2, 1, and 0.5 that correspond to the observed soft and hard bands at z = 0, as
well as the rest-frame B-band and FIR band.
-- 21 --
Fig. 3. -- Potential influence of obscuration on multi-wavelength QLFs at z ∼ 2. Solid lines show
B-band, FIR, 0.5-2 keV, and 2-10 keV luminosity functions of a model in which all quasars have
the Elvis et al. (1994) SED, with p( m) and n(M ) chosen to match the Boyle et al. (2000) B-band
results (triangles in upper left). Dashed lines show QLFs for a model in which 20% of quasars have
the Elvis et al. (1994) SED and 80% have an obscured SED corresponding to NH = 3 × 1023cm−2
(see Fν values in Table 1). Dotted lines show the contribution of obscured systems in this model
(note that we assume complete optical obscuration, hence no contribution in B). X-ray luminosities
are observed-frame at z = 2.
-- 22 --
of obscured quasars (dotted line) dominating at all luminosities. For the 0.5-2 keV band (rest-
frame 1.5-6 keV) the situation is more complicated, since obscuration suppresses flux in this band
by nearly a factor of ten. At high luminosities, unobscured quasars dominate the QLF because
the greater numbers of obscured quasars are not enough to compensate for their reduced fluxes.
However, the obscured population boosts the faint end of the soft X-ray QLF by about a factor of
two, with obscured and unobscured systems making roughly equal contributions.
The most dramatic effect of the obscured population is to boost the FIR luminosity function
by a large factor, since with our assumptions the FIR band contains 79% of the bolometric flux of
obscured systems but only 17% of the bolometric flux of unobscured systems. The combination of
more systems and more flux per system boosts Φ(LFIR) by almost two orders of magnitude at high
luminosities, with the FIR QLF totally dominated by obscured systems at every redshift. Note that
our treatment of FFIR implicitly assumes that obscured and unobscured systems are two distinct
populations, one with high gas columns and one without. It is also possible, as assumed by Sazonov,
Ostriker, & Sunyaev (2003), that the difference between obscured and unobscured systems is simply
one of orientation, and that even systems that are unobscured along our line of sight have most of
their optical, UV, and soft X-ray emission absorbed by a dusty torus and re-radiated isotropically
in the FIR. In this case, FFIR would be essentially the same for both populations, so at a given
FIR luminosity obscured and unobscured systems would be represented in their global ratio (i.e.,
4:1 in our model). The joint FIR-optical or FIR-soft X-ray luminosity functions would distinguish
these two scenarios.
Figure 3 shows that a large population of obscured quasars can substantially alter the relation
between B-band, X-ray, and FIR luminosity functions, as one would expect. The difference in
FIR would persist at all redshifts. At low redshifts, on the other hand, the 0.5-2 keV band is
almost completely suppressed by a column density NH = 3 × 1023cm−2, and the 2-10 keV band is
significantly suppressed, so the effect of an obscured population on the luminosity function in these
bands is reduced.
Although we include the ADAF mode in our calculations for Figure 3, we find that ADAF
systems make no significant contribution to the QLF at any luminosity likely to be observed at
z ∼ 2, for any plausible choice of our model parameters. However, the situation could be different
at low redshift, in part because observations reach to lower luminosities, but even more because (as
we discuss in §4.2.1 below) matching the observed QLF evolution requires a shift of p( m) towards
lower characteristic accretion rates at low redshifts, thus giving systems with m < mcrit more
chance to compete. Figure 4 shows two models with choices of p( m) and n(M ) that give reasonable
matches to the Boyle et al. (2000) B-band QLF at z ∼ 0.5. The upper panel shows the two p( m)
distributions, which are double power-laws with m∗ = 0.03 and 0.012, respectively. The black hole
mass functions (not shown) have corresponding M∗ values of 1.21 × 109M⊙ and 2.5 × 109M⊙.
Solid lines in the middle and bottom panels show the predicted luminosity functions for the first
model in rest-frame B-band and observed-frame 2−10 keV, respectively. Total luminosity functions
-- 23 --
Fig. 4. -- Potential influence of ADAFs on multi-wavelength QLFs at z ∼ 0.5. The top panel shows
p( m) distributions for two models with different m∗ (the values of M∗ are also different). The total
QLFs for the two models are nearly identical in both B-band (middle panel) and 2-10 keV (bottom
panel); solid lines show the total QLFs of Model 1. Dotted and dashed lines show the contributions
of ADAF accretion ( m < 0.01) for the two models, which are small in B-band but substantial at
2-10 keV for Model 2.
-- 24 --
for the second model are nearly identical. However, the relative contribution of the ADAF mode,
shown by the dotted and dashed lines in these panels, is quite different between the two models and
between the two bands. In B-band, ADAF contributions to the luminosity function are strongly
suppressed, even for m∗ = 0.012, because of the low optical flux of the ADAF SED. However, the
relatively high hard X-ray flux allows ADAF accretion to dominate the low end of the luminosity
function for the m∗ = 0.012 model and to make a significant contribution even at high luminosities.
In the m∗ = 0.03 model, on the other hand, ADAF accretion is a subdominant contribution to the
QLF at all luminosities. Results for the FIR and 0.5 − 2 keV luminosity functions are qualitatively
similar to those for B-band and 2 − 10 keV, respectively, as one would expect from the values
of Fν in Table 1. Measurements of the optical and X-ray luminosity functions alone would not
distinguish the two models shown in Figure 4, but the low- m∗ model predicts that X-ray selected
quasars (largely ADAF systems) should have systematically different SED shapes from optically
selected quasars (mostly thin-disk systems), while the high- m∗ model predicts that thin-disk SEDs
dominate both populations.
4. Evolution
We now turn to evolutionary calculations, applying the basic principles of §2.2. We assume that
the accretion physics -- the dependence of ǫ0.1 and SED shape on accretion rate -- is independent
of redshift, although the distribution of accretion rates itself evolves. This assumption seems
reasonable, since the "microphysics" has no direct knowledge of the age of the universe, but one
could imagine that systematic changes in the host galaxy population might affect the influence of gas
or dust obscuration on SED shapes, and perhaps even that galaxy mergers could alter the fraction
of spinning black holes with higher bolometric efficiencies. With our assumption and a specified
model of the accretion physics, p( mM, z) determines the evolution of Φ(L) in all wavebands, since
it both determines the evolution of n(M ) and specifies the probability that black holes of a given
mass shine at a given luminosity. However, mergers can alter the evolution of the QLF by changing
n(M ) independently of p( mM, z).
We focus in this section on the optical luminosity function, which is observed to rise by a
large factor between z ∼ 5 and z ∼ 3 (Warren, Hewett, & Osmer 1994; Schmidt, Schneider, &
Gunn 1995; Fan et al. 2001) and decline by a large factor between z ∼ 2 and z ∼ 0 (Schmidt
1968; Boyle et al. 2000). At z . 2.5, Boyle et al. (2000) find a break in the luminosity function
(eq. 19) that evolves towards lower luminosities at lower redshifts, a form of "luminosity evolution"
that cannot be described by a simple vertical shift in amplitude ("density evolution"). We will
devote considerable attention to the implications of this result, though we should note that Wolf
et al. (2003) reach a more ambiguous conclusion about the need for luminosity evolution, using a
data set (COMBO-17) that reaches still lower luminosities. At z & 2.5, current observations probe
mainly the high luminosity end of the QLF, where the data are adequately described by a single
power-law.
-- 25 --
We begin our discussion below with a few remarks on the definitions of quasar lifetimes and
duty cycles. We then present evolutionary calculations for a number of specific models designed to
illustrate general points, working within the class of models discussed in §3.2: double power-law
p( m), double power-law n(M ), and ǫ0.1( m) as defined in equation (7). In §4.2, we consider cases
in which accretion alone drives the evolution of n(M ), looking first at the declining phase of QLF
evolution (z ≤ 2) and then at the growing phase. In §4.3, we use simple models to illustrate the
potential impact of mergers on the evolution of n(M ) and the QLF.
4.1. Lifetimes and duty cycles
One of the key elements in models of quasar evolution is the typical quasar lifetime, or, nearly
equivalent, the duty cycle of quasar activity (see Martini [2003] for a review of observational esti-
mates). In a simple "on-off" model of the quasar population, where a black hole is either shining at
a fixed fraction of its Eddington luminosity or not accreting at all, it is clear what these concepts
refer to: at a given redshift, the duty cycle is the fraction of black holes that are active at any one
time, and the typical lifetime is the integral of the duty cycle over the age of the universe. However,
if p( m) is broad, and in particular if there is a tail of increasing probability towards low m, then
defining a black hole to be "on" if it is accreting at any non-zero rate may not be particularly
useful, since changes to p( m) that have negligible observational effect (such as changing the lower
cutoff mmin) may have a large impact on the implied duty cycle or lifetime. Such a definition is
also difficult to relate to black hole growth or emissivity.
For our purposes, it is more useful to consider the accretion weighted lifetime
tacc ≡Z tf
ti
m(t) dt .
(23)
The ratio of tacc to the Salpeter lifetime (ts = 4.5 × 107 yr, cf. eq. 5) gives the number of e-folds of
mass growth, i.e.,
ts (cid:21) .
Mf = Mi exp(cid:20)tacc
(24)
Typically, the initial time ti would refer to some time after the formation of "seed" black holes but
before the main epoch of mass accretion, and tf would refer to z = 0. However, since we model
the declining and growing phases of quasar evolution separately below, we will generally choose ti
and tf to correspond either to the redshift interval z = 2 − 0 or to the redshift interval z = 5 − 2.
A useful way to characterize the mean accretion rate at a given redshift is
tacc,z ≡ h miH −1(z) .
(25)
The ratio tacc,z/ts is the number of e-folds of mass growth that would occur if the mean accretion
rate were to stay constant for the Hubble time H −1(z).
-- 26 --
For constant efficiency ǫ0.1, weighting the lifetime by m is equivalent to weighting by L/(ǫ0.1LEdd),
so the accretion weighted lifetime is simply related to the luminosity weighted lifetime. Alterna-
tively, one can weight activity by the ratio of a black hole's luminosity L(t) to its final Eddington
luminosity lMf . In this case, the weighted lifetime (for constant efficiency and no black hole merg-
ers) is tsǫ0.1(1 − Mi/Mf ), which approaches the Salpeter lifetime in the limit that the black hole
mass grows by a large factor.
4.2. Pure Accretion Driven Evolution
4.2.1. Declining Phase with Mass Independent p( m)
For pure accretion driven evolution, only the first term of equation (12) enters into the evolution
of n(M, t). We start by considering the case with p( m) independent of mass, so that the "self-
similar" solution given in equation (11) applies. For a double power-law p( m), the evolution of the
QLF depends on the time evolution of the normalization p∗(t) and the characteristic accretion rate
m∗(t). We initially consider the declining phase of QLF evolution and take both functions to be
power-laws of time:
p( mt) =
p∗(t)(cid:16) m
p∗(t)(cid:16) m
m∗(t)(cid:17)a
m∗(t)(cid:17)b
m < m∗(t)
m > m∗(t)
,
p∗(t) = p∗,i(cid:18) t
ti(cid:19)γp
,
m∗(t) = m∗,i(cid:18) t
ti(cid:19)γm
, (26)
where ti represents the time from which the initial QLF is evolved, and p∗,i and m∗,i correspond to
the values of p∗ and m∗ at time t = ti.
For γm = 0, "pure p∗ evolution," the relative probability of given accretion rates remains fixed
over time, but the overall probability of accreting per unit time declines. This is analogous to "pure
density evolution" models of the QLF, though the masses of the black holes continue to evolve.
For γp = 0, "pure m∗ evolution," the relative probabilities of given accretion rates change over
time as the break in p( m) evolves. This is analogous to "pure luminosity evolution," though again,
the black hole mass function is evolving along with the evolution of p( m). Physically, p∗ evolution
could be connected to a decline in the frequency of galaxy interactions as the universe gets older.
Evolution of
m∗ could arise from declining gas fractions of galaxies, or a decline in their ability to
funnel gas to the center as they become larger and more stable.
The evolution of the black hole mass function can be determined by using equation (11) to
express n(M, t) as
n(M, t) =
n∗(cid:16) M
n∗(cid:16) M
M∗(t)(cid:17)α M∗,i
M∗(t)(cid:17)β M∗,i
M∗(t) M < M∗(t) ,
M∗(t) M > M∗(t) ,
M∗(t) = M∗,i exp(cid:18)Z t
ti
h m(t)i
dt
ts(cid:19) ,
(27)
where M∗,i is the mass corresponding to the break in the black hole mass function at a time
t = ti. Since we assume that p( m) is independent of mass, the slopes of the QLF at low and high
-- 27 --
luminosities match the slopes of n(M ), and thus we choose the Boyle et al. (2000) slopes α = −1.5
and β = −3.4 for n(M ). We use M∗ = 109M⊙ as in §3.2. For p( m), the values a = −0.5, b = −3.0,
m∗,i = 0.9 then give a reasonable fit to the data at z = 2. For our choice of p( mt), the integral for
M∗(t) can be done analytically.
Figure 5 uses equations (26) and (27) to determine the black hole mass function and probability
function at redshifts z = 2, 1, and 0.5. The left hand panels show M∗(t) in the lower windows and
p( mt) in the upper windows. The right hand panels show the evolution of the QLF generated by
the functions in the corresponding left hand panels, using the methods and assumptions described
in §3.2. To reduce the dimensionality of the parameter space, we assume a value of γm and then
find a value for γp that makes the evolved, z ∼ 0.5 QLF match the Boyle et al. (2000) data at the
break luminosity Lbrk. As a characterization of the mean accretion level, we list in the left hand
panels the values of the local accretion weighted lifetimes tacc,z (eq. 25), in units of the Salpeter
time ts, at z = 2, 1, and 0.5. These ratios tacc,z/ts give the number of e-folds of black hole growth
that would occur in a Hubble time H −1(z) if p( m) stayed constant. In contrast to earlier figures,
we plot m2p( m) rather than p( m) itself, because this product gives the contribution to black hole
growth (and emissivity of the quasar population) per logarithmic interval of
m. For our adopted
forms of p( m), the largest contribution to growth and emissivity always comes from accretion rates
near m∗.
The upper panels of Figure 5 show an example with γm = 0, so that the normalization of
p( mt) evolves but the shape does not. Although we require the predicted QLF to pass through
the observed Φ(Lbrk) at each redshift, the shapes of the model luminosity functions at z = 1 and
z = 0.5 disagree strongly with the data. Generally, evolution of p∗ alone cannot reproduce the type
of QLF evolution found by Boyle et al. (2000) at low redshift, because the growth of M∗ combined
with a fixed relative distribution of accretion rates shifts the predicted QLF horizontally to a higher
break luminosity, while the observed break luminosity declines with time. Reducing p∗ reduces the
fraction of accreting black holes, but that only produces a vertical drop of the QLF, which cannot
fully compensate for the shift to a higher break luminosity.
Since "pure density evolution" models fail to describe the Boyle et al. (2000) data, it is no
surprise that our analogous "pure p∗ evolution" models also fail. However, Figure 5 shows that
black hole growth exacerbates the failures of a model in which only the frequency of fueling activity
declines with time, since such models generically predict an increase in the break luminosity with
time. We conclude that, if p( m) is independent of mass, then it must evolve in a way that increases
the relative probablilities of low accretion rates in order to make the QLF evolve to lower break
luminosities. We will consider an alternative explanation -- that p( mM ) has a mass dependence
that evolves with time -- in §4.2.2. However, a decline in typical values of m does seem a plausible
consequence of decreasing gas fractions and increasing stability of host galaxies, and such changes
in galaxy properties are likely to play an important role in producing the observed form of QLF
evolution. Furthermore, even if gas fueling rates remain fixed in physical units, they decline in
Eddington units as black hole masses increase, thus driving m values down if black holes grow
-- 28 --
Fig. 5. -- Three models of the declining phase of QLF evolution, from z = 2 to z = 0.5. For each
model, left hand panels show m2p( m) at z = 2, 1, and 0.5 in the upper window (solid, dotted, and
dashed lines, respectively), and M∗(z) in the lower window. Each panel also lists the accretion
lifetime tacc,z = h miH −1(z) in units of the Salpeter timescale ts = 4.5 × 107 yr, for z = 2, 1, and 0.5
(top to bottom). Right hand panels show the corresponding QLFs at the three redshifts. Points
show the fits of Boyle et al. (2000) at z = 2 (triangles), z = 1 (squares), and z = 0.5 (circles)
over roughly the absolute magnitude range probed by their data. The top panels show a model in
which m∗ stays fixed and only the normalization p∗ declines with time. The middle panels show
a model with the same initial p( m) but declining evolution of
m∗. The bottom panels show a
similar model that starts with a higher black hole space density and lower normalization of p( m),
and consequently less black hole growth. Matching the observed evolution of Lbrk towards lower
luminosity requires a decrease in the characteristic accretion rate m∗, not merely a decrease in duty
cycle.
-- 29 --
by a significant factor.
In their semi-analytic model of the quasar and host galaxy population,
Kauffmann & Haehnelt (2000) find that they must account for the decreasing gas supplies and
longer dynamical timescales of host galaxies at low redshift, in addition to the decreasing frequency
of mergers, to explain the observed evolution of the QLF. In terms of our models, the first two
effects are analogous to decreases in m∗, while the last is analogous to a decrease in p∗.
The models in the middle and bottom rows of Figure 5 incorporate declining m∗ and match
the observed QLF evolution better. The case in the middle row starts with the same p( m) and
n(M ) at z = 2, but it has γm = −2.7, which moves the break in p( mt) to lower accretion rates as
time progresses. Though the match to the data is not perfect, it is much better than before, with
the break in Φ(L) shifting to lower luminosities as the QLF becomes increasingly dominated by
the contribution from lower accretion rates. The model in the bottom row starts at z = 2 with a
black hole space density a factor of ten higher and an accretion duty cycle a factor of ten lower (n∗
and p∗ increased and decreased by ten, respectively). The QLFs at z = 2 are identical because of
the exact degeneracy between n∗ and p∗ discussed in §3.1. However, the reduction in p∗ lowers the
mean accretion rate h mi, which in turn leads to less black hole growth: this model has tacc,z < ts
at all z < 2, and the characteristic mass M∗ hardly grows at all. The shift to lower accretion
rates coupled with the smaller amount of black hole growth over time yields QLF evolution in good
agreement with the data.
Figure 5 shows that the degeneracy between n∗ and p∗ is broken once evolution is taken into
account. If the black hole density is low, then each black hole must accrete more in order to match
the observed QLF, and this accretion leads to more rapid evolution of n(M ). In the case shown in
Figure 5, the model with less black hole growth matches the data better. However, it is possible
to start with the initial conditions of the model in the middle panels and match the data nearly
as well by dropping m∗ more rapidly, with γm = −3.7 instead of −2.7. Thus, changes to n∗ and
p∗ can be partly compensated by changes to other parameters. Nonetheless, evolution narrows the
range of the n∗p∗ degeneracy because models with too much black hole growth (too low n∗) cannot
yield a declining break luminosity for any plausible evolution of p( m). Furthermore, as discussed in
§3.3 and further in §5 below, models with different n∗ predict different distributions of active black
hole masses and accretion rates, different space densities of host galaxies, and of course different
underlying n(M ), even when they match the same, evolving QLF.
Similar remarks apply to the partial degeneracy between M∗ and m∗ discussed in §3.2. Com-
binations of M∗ and m∗ that yield similar QLFs at z ∼ 2 will have more growth of n(M ), and thus
different evolution, if M∗ is lower and h mi consequently higher.
4.2.2. Declining Phase with Mass-Dependent p( m)
As shown in §3.4, mass dependence of p( m) can break the link between the shape of the
black hole mass function and the shape of the luminosity function, adding considerable freedom
-- 30 --
to models of the QLF. In an evolutionary calculation, one must also account for the influence of
mass-dependent growth on the shape of the black hole mass function. If the more numerous, low
mass black holes have a higher probability of being active, then they grow faster than the high
mass black holes, and the mass function steepens. Conversely, if high mass black holes are more
active, then they grow faster and the mass function becomes shallower with time. It is intuitively
useful to think of this behavior in graphical terms. On a log-log plot, a mass-independent p( m)
causes the mass function to shift horizontally in a coherent fashion, maintaining its shape. Faster
growth of low mass black holes allows the low end of n(M ) to translate faster and "catch up" with
the high end. Conversely, faster growth of high mass black holes allows the high end of n(M ) to
stretch away from the low end. As always (eq. 10), it is only the mean accretion rate h m(M )i
that matters for determining the evolution of n(M ), and one can calculate the evolution exactly
by assuming that all black holes of a given mass accrete at this rate.
To describe the evolution of n(M ) mathematically, we define g(M, t) = Mi/M , where Mi and
M represent black hole masses at times ti and t, respectively. Matching number densities in the
equivalent mass intervals at the two times then implies that n(M )dM = ni(Mi)dMi, and thus
n(M ) = ni(Mi)
dMi
dM
= ni(Mi)(cid:20)g(M, t) + M
∂g
∂M(cid:21) .
(28)
If g(M, t) is independent of mass, then we have simple remapping of Mi → M and renormalization
by g(M, t), recovering the result (11). For mass-dependent h mi, the value of n(M ) at mass M may
be higher or lower than the mass-independent case depending on the sign of ∂g/∂M . Equation (28)
allows the numerical calculation of n(M ) for any specified p( mM, z), since this determines g(M, t).
However, the exponential relation between the growth factor andR h midt generally means that any
simple analytic form of n(M ) is lost once black holes grow significantly.
We have previously adopted mass-independent p( m) as a mathematical convenience, but con-
sideration of black hole growth suggests that this choice is not completely arbitrary. Suppose that
the black holes in some mass range have a high h mi relative to their peers because they tend to
reside in galaxy hosts that can feed them more efficiently. These black holes e-fold in mass more
rapidly than others, and their fueling rates in Eddington units therefore drop more quickly (or
grow more slowly), bringing them back into line. This regulating mechanism suppresses mass-
dependence of h mi, causing n(M ) to change shape until it approaches the "fixed point" solution
of mass-independent h mi, after which it evolves in a self-similar fashion. Constancy of h mi does
not necessarily imply constancy of p( m), but this regulation argument suggests that p( m) might be
approximately mass-independent in an average sense at redshifts near the peak of quasar activity.
The regulating mechanism may lose force at low redshift, when black holes no longer grow by
substantial factors and gas merely trickles onto full grown systems. Thus, we might expect stronger
mass dependence of p( m) in the declining phase of quasar evolution. We have shown in §4.2.1 that
matching the observed shift of Lbrk to lower luminosity requires a decline in characteristic m values
if p( m) is independent of mass. However, mass-dependent p( m) offers another possibility: activity
-- 31 --
could decline preferentially in more massive black holes between z = 2 and z = 0, thus driving the
break luminosity down as the typical mass of active systems declines.
To create a model along these lines, it is helpful first to consider the case where there is no
evolution of n(M ) at all, so that one can infer the required p( mM, z) from a simple graphical
argument relating the vertical shift of Φ(L) to the horizontal shift of Lbrk. We consider a double
power-law QLF with slopes α and β below and above Lbrk respectively, and we assume pure
luminosity evolution with Φ(Lbrk) = constant and Lbrk(t2) < Lbrk(t1) for t2 > t1. For maximum
contrast with the model in §4.2.1, we assume that the mass dependence of p( m) enters only in the
normalization p∗(M ), not in the slopes or in the characteristic accretion rate m∗. In other words,
the relative distribution of accretion rates remains the same at all times for all active black holes,
but the probability of a black hole being active at all depends on its mass, in a redshift-dependent
manner. Relating the amplitudes of Φ(L) at times t1 and t2 to the horizontal shift of Lbrk from
time t1 to t2 then implies
Φ(L, t2)
Φ(L, t1)
=( kα L < Lbrk(t2)
kβ L > Lbrk(t1)
,
k =
Lbrk(t2)
Lbrk(t1)
,
(29)
with a more complicated dependence in the range Lbrk(t2) < L < Lbrk(t1). Since n(M ) is constant,
any shift in the QLF must be produced by a shift in p∗(M ), and thus,
p∗(M, t2)
p∗(M, t1)
kα
kβ
kα + (kβ − kα)U (M ) Mmin < M < Mmax ,
M < Mmin
M > Mmax
(30)
=
where Mmin is the minimum mass of a black hole that can generate L = Lbrk(t2) and Mmax is
the maximum mass of a black hole that can generate L = Lbrk(t1). Here U (M ) is some function
that goes from zero to one as mass goes from Mmin to Mmax, and it can be tuned to reproduce
the observed Φ(L) in the break region. With equation (30) as a starting guess, we can find by
numerical iteration a solution with mass-dependent p∗(M ) that reproduces the observed evolution
and self-consistently incorporates the implied growth in n(M ).
Figure 6 shows a model in which we assume that p( m) is independent of mass at z = 2 and that
subsequent evolution of the QLF (left panel) is driven by the mass dependence of p( mM ) (right
panel). We choose a normalization of the black hole mass function, n∗M∗ = 1.2 × 10−4 Mpc−3,
that corresponds to a fairly short quasar lifetime, tacc = 7 × 106 yr (from z = 2 to z = 0), so
that the mass-dependent growth does not severely distort the double power-law form of n(M ) that
our initial guess at p( mM ) assumes. This model fits the Boyle et al. (2000) data as well as the
model with mass-independent p( m) shown in the bottom panel of Figure 5. With regard to Φ(L)
alone, the two models are effectively degenerate, but Lbrk in the mass-dependent model evolves to
lower luminosities because high mass black holes are less likely to be active at low redshift. Thus,
relative to the mass-independent model, this model predicts that luminous AGN at low redshift
consist primarily of low mass black holes with high L/LEdd, and it predicts a narrower range of
m
-- 32 --
Fig. 6. -- An alternative model for the declining evolution of the QLF, in which the characteristic
accretion rate m∗ remains constant but the mass dependence of p( mM ) evolves to suppress activity
preferentially in high mass black holes at lower redshifts. As in Figure 5, triangles, squares, and
circles in the left hand panel represent the Boyle et al. (2000) observational fits at z = 2, 1, and
0.5, respectively, while lines show the model predictions (including an extrapolation to z = 0).
The right hand panel shows the evolving mass dependence of the normalization p∗(M ), which is
proportional to the duty cycle of black holes of mass M . The model assumes our usual double
power law forms of n(M ) and p( m), with M∗ = 109M⊙ and a mass-independent p( m) at z = 2,
and m∗ = 0.5 at all redshifts.
-- 33 --
values. We compare the two models' predictions for the mass distribution of active systems in §5
below (Fig. 16).
4.2.3. Growing Phase
We now turn to the redshift interval z ∼ 5 to z ∼ 2, during which Φ(L) first grows rapidly, then
reaches a plateau between z ∼ 3 and z ∼ 2 (see, e.g., Warren, Hewett, & Osmer 1994; Schmidt,
Schneider, & Gunn 1995; Pei 1995). In the range z = 3.6 − 5.0, Fan et al. (2001) provide the best
measurements of the bright end of the luminosity function, while Wolf et al. (2003) give constraints
at fainter luminosities. To cover the gap between z = 3.6 and the Boyle et al. (2000) measurements
at z = 2, we use the measurements of Warren, Hewett, & Osmer (1994), with a median redshift
z ≈ 3.25. We use the Ωm = 1 cosmological model, since all of these papers give results for this case,
and we adopt h = 0.5 so that the age of the universe is realistic. We convert the Warren, Hewett,
& Osmer (1994) space densities and absolute magnitudes from h = 0.75 to h = 0.5, and we convert
their AB(λ1216A) magnitudes to B-band magnitudes using MB =AB(λ1216A)−0.605, assuming
fν ∝ ν −0.5 for the conversion of AB(λ4400A) to AB(λ1216A). Our goal here is not to model the
data in detail but merely to illustrate what kinds of p( m) and accompanying n(M ) evolution can
fit the general trends.
For simplicity, we restrict ourselves to the class of models in which p( m) is a double power-law
independent of mass and the characteristic accretion rate m∗ is constant from z = 5 to z = 2.
At z = 2, we assume a double power-law n(M ) with M∗ = 109M⊙ and slopes α = −1.5 and
β = −3.4. The evolution of the QLF is then determined by the evolution of the amplitude of
p( m), which we assume has a piecewise power-law form, p∗(t) ∝ tγp in the interval between each
of the redshifts where we match the QLF, though we allow γp to be different from one interval to
another. Our general procedure is to take a model that fits the QLF data at z = 2, then evolve it
to higher redshift, iteratively solving for values of γp in each redshift interval so as to match the
observed amplitude of the QLF at MB ∼ −26.5, using the Warren, Hewett, & Osmer (1994) data
at z = 3.25 and the Fan et al. (2001) data at z = 3.75, 4.15, and 4.7. Iteration is necessary because
we must calculate the time integrated mean accretion rate for the given γp and reduce black hole
masses by the corresponding factor, before calculating the QLF with the evolved p( m). Because
mass-independent p( m) leads to self-similar evolution of n(M ) and identical asymptotic slopes for
n(M ) and Φ(L) (see §3.1), this restricted class of models cannot explain the apparent change in
the bright-end slope of the QLF (see Fan et al. 2001) between z ∼ 4 and z ∼ 2.
Figure 7 shows two qualitatively different solutions that reproduce the observed evolution
of the QLF amplitude. The first of these solutions, shown in the top panels, has a relatively
high normalization of the black hole mass function at z = 2, with space density n∗M∗ = 1.2 ×
10−4 Mpc−3. The high space density leads to a low normalization of p( m) at each redshift, and
consequently to little growth of black hole masses; the lifetimes tacc,z are shorter than ts at all
redshifts, and the value of M∗ increases only slightly over the entire range z = 4.7 to z = 0. (Note
-- 34 --
Fig. 7. -- Like Figure 5, but for the growing phase of QLF evolution from z ∼ 5 to z ∼ 2. Points in
the right hand panels are based on Fan et al. (2001) at z ∼ 4.7, 4.15, and 3.75 (pentagons, squares,
filled triangles), on Warren, Hewett, & Osmer (1994) at z ∼ 3.25 (circles), and on Boyle et al.
(2000) at z ∼ 2 (open triangles). Model parameters are chosen to reproduce the amplitude of the
observed QLF at MB ∼ −26.5. The top row shows a model with a high black hole space density at
z = 2, n∗M∗ = 1.2 × 10−4 Mpc−3, in which case there is little growth of black hole masses between
z = 5 and z = 2 and the QLF evolution is driven by increasing p( m). The bottom row shows a
model with a lower black hole space density and, therefore, more accretion per black hole. In this
model, the growth of the QLF is driven mainly by the increasing masses of black holes.
-- 35 --
that we show the low redshift evolution of M∗ inferred by continuing the model past z = 2 using
the Boyle et al. [2000] data as in the previous section.) The growth in the amplitude of the QLF
is therefore driven by the steadily rising amplitude of p( m), with no significant contribution from
growth in n(M ) itself.
While this high space density solution fits the QLF data reasonably well by construction, it
seems rather implausible on physical grounds. The minimal evolution of M∗ means that all of the
observed quasar activity represents a negligible contribution to the growth of black holes. Instead,
all of the supermassive black holes must have been assembled in some unseen manner before z = 5,
and the observed evolution of the QLF reflects a gradual "turning on" of these black holes at lower
redshift. In such a model, the "seed" black holes formed at z > 5 are essentially the same black
holes that are present today, though mergers may have converted many low mass black holes into
a smaller number of high mass systems.
The bottom panels show a solution near the opposite extreme, with a lower space density
n∗M∗ = 2.012 × 10−5 Mpc−3 at z = 2. Matching the QLF now requires a higher normalization
of p( m) at each redshift, and the accretion weighted lifetimes are in the range tacc,z ∼ 0.4 − 2.7ts.
There is roughly a factor of ten growth in M∗ between z = 5 and z = 2, and this growth plays a
central role in the evolution of the QLF. In contrast to the high space density case, the amplitude
of p( m) is approximately constant from z = 5 to z = 3, with a slight drop to z = 2.
In this
solution, therefore, the observed quasar activity traces the growth of the black hole population
from much smaller seeds present at z = 5, with a roughly constant distribution of accretion rates
in Eddington units during the growing phase of quasar evolution. The break in the QLF moves
to higher luminosity as M∗ grows, so measurements probing to lower luminosities at high redshift
could distinguish this solution from the short lifetime solution shown in the top panels.
sity ρbh =R ∞
If the n(M ) normalization is reduced just slightly further, below n∗M∗ ≈ 2.01 × 10−5 Mpc−3,
then the solution for the evolving n(M ) requires unphysical, negative densities at z = 5. Thus,
within our assumptions, there is a minimum allowed space density of black holes, which corre-
sponds to the limit in which the observed QLF traces all of the accretion onto the black hole
population. We can see why this is so by recalling Soltan's (1982) argument that the integrated
bolometric emissivity of the quasar population determines, for an assumed efficiency, the mass den-
0 M n(M )dM of the black hole population. We have adopted a form for n(M ), and the
evolutionary calculation allows us to, in effect, correct the observed emissivity for the contribution
from lower luminosity systems. Since we have also specified the efficiency ǫ( m) and the bolometric
correction Lbol/LB, matching the observational data points determines ρbh at z = 2 and thereby
fixes the normalization of n(M ). Our results thus show that a suitable extension of the Soltan
(1982) argument, aided by some auxiliary assumptions and a measurement of the QLF at z ∼ 2,
can predict the black hole mass function n(M ) itself, not just the integral ρbh. We will explore this
idea in future work, with attention to the sensitivity of the predictions to the auxiliary assumptions
and to uncertainties in the observational data.
-- 36 --
4.3. Mergers
The general equation (12) for the evolution of the black hole mass function has a very limited
set of analytic solutions, even if one considers only the merger terms and ignores accretion driven
growth. Realistic calculations incorporating merger driven growth will probably need to be done
numerically, with some a priori model (based on galaxy merger trees, for example) for what merger
rates should be. Here we will investigate some simple, analytically solvable cases that can provide
insight into the generic effects of mergers on the black hole mass function.
First, we consider a binary merger model in which there is a probability f per unit time that
a given black hole merges with another black hole of equal mass to form a new black hole of twice
the original mass. In the absence of accretion, a counting argument yields the equation governing
n(M, t),
∂n(M, t)
∂t
= −f n(M, t) +
1
4
f n(cid:18)M
2
, t(cid:19) .
(31)
The first term on the r.h.s. is the sink representing loss of black holes of mass M to mergers, and
the second is the source representing creation by mergers of systems with mass M/2, with a factor
of 1/4 to account for the replacement of two black holes by one and the factor of two growth in
the dM interval. The only solution to (31) in which n(M ) maintains its shape over time is a pure
power-law of the form
M∗(cid:19)α
n(M, t) = n∗(t)(cid:18) M
,
(32)
where we have included all the time dependence in n∗ because the effects of n∗ and M∗ are degen-
erate for a pure power-law. With this form, the solution to (31) is
n∗(t) = n∗(ti) exp(cid:20)Z t
ti (cid:16)2−(α+2) − 1(cid:17) f (t)dt(cid:21) .
(33)
If f (t) is constant, then n∗(t) evolves exponentially in time, while f (t) ∝ t−1 yields n∗(t) evolving
as a power-law with slope of (cid:0)2−(α+2) − 1(cid:1) tif (ti).
The important general feature of the solution (33) is that the sign of the evolution depends on
the slope α of the mass function. For the critical value α = −2, mergers do not change n(M ) at all,
because the source and sink terms in equation (31) balance. If α < −2 then n(M ) increases with
time, and if α > −2 then n(M ) decreases with time. For a steep n(M ), the black holes added to a
given range of mass from mergers of lower mass objects exceed the number lost to higher masses.
For a shallow n(M ), on the other hand, mergers consume more black holes in a given mass range
than they create. While the solution (33) is specific to our restricted model, different behavior for
steep and shallow slopes of n(M ) follows from mass conservation, so we expect it to hold quite
generally.
The pure power-law n(M ) adopted above cannot hold for all masses because the implied total
mass density of black holes would be infinite. However, the behavior of the single power-law solution
-- 37 --
gives insight into the more general case of a mass function that changes slope from low to high
masses. For example, a double power-law has a steep high mass end where mergers drive n(M ) up
with time and a shallow low mass end where mergers drive n(M ) down. Over the range in mass
near the break, the shape changes as the number of high mass black holes grows and low mass
black holes decreases, making the break smoother and shifting it to lower masses. Again, we expect
these effects of mergers to be fairly generic.
For a pure power-law n(M ), we can also include the accretion term of equation (12) in the
calculation, obtaining the solution
n
n
=
n∗
n∗
= −
h m(t)i
ts
(1 + α) +(cid:16)2−(α+2) − 1(cid:17) f (t) .
(34)
The sign of the accretion term also depends on α, but here the critical slope is −1 instead of
−2, reflecting the fact that accretion adds mass to the black hole population while mergers do
not. Equation (34) can be integrated analytically if accretion and merger rates have same time
dependence, h m(t)i = h m(ti)ih(t) and f (t) = f (ti)h(t), with h(t) an arbitrary function having
h(ti) = 1. The solution is
n∗(t) = n∗(ti) exp(cid:18)(cid:20)−
h m(ti)i
ts
(1 + α) +(cid:16)2−(α+2) − 1(cid:17) f (ti)(cid:21)Z t
ti
h(t)dt(cid:19) .
(35)
Up to factors that are typically of order unity, the relative importance of accretion and mergers
depends on the value of h m(ti)i relative to tsf (t), the average number of mergers per Salpeter time.
However, the pre-factors can greatly diminish one term or the other close to the critical slopes
α = −1 or α = −2, and for −1 > α > −2 accretion and mergers affect n(M ) in opposite directions.
For our illustrative model calculations in §5, we do not want to assume a pure power-law
n(M ), and we will therefore adopt another very simple prescription for mergers, similar to that of
Richstone et al. (1998). We assume that accretion has negligible impact on n(M ) after some time
t1, which is true if h m(t1)it1 ≪ ts and h m(t)i falls as t−1 or faster. At some later time, t2, every
black hole with initial mass M1 is assumed to have merged with (fm − 1) other black holes of mass
M1 to make one black hole of mass M2 = fmM1. The black hole mass functions at t1 and t2 are
related by the transformation
n2(M2)dM2 =
1
fm
n1(cid:18)M1 =
M2
fm(cid:19) dM1
fm
or simply
n2(M ) = f −2
m n1(M/fm).
,
(36)
(37)
In this model as in our previous model, mergers drive n(M ) up when the logarithmic slope is
α < −2 and down when α > −2. If n(M ) is a double power-law at time t1, then at time t2 it is still
a double power-law, with M∗ larger by a factor fm and the normalization lower by a factor f 2
m.
In both of our merger models, we assume that mergers conserve the total mass of the black
hole population, and simply redistribute it from low mass systems to high mass systems. However,
-- 38 --
it is also possible for mergers to decrease the total mass of the population, at least those black holes
that reside at the centers of galaxies and have the potential to become quasars. This can happen
if multiple mergers produce triple or quadruple systems that lead to ejection of one or more of
the black holes from the galaxy (e.g., Valtonen et al. 1994). Even a merging binary system can
potentially be ejected by a gravitational radiation "rocket" effect, though this seems more likely to
be important in shallow potential wells hosting low mass black holes (see Redmount & Rees 1989;
Madau et al. 2003). Finally, a binary could remain at the galaxy center but radiate a significant
fraction of the mass of its progenitors in gravity waves during the merger event (Yu & Tremaine
2002 and references therein). We will not consider any of these possibilities in detail here, but we
note that in all these cases the critical slope at which n(M ) grows rather than declines would be
steeper than −2, since a given bin would lose mass to mergers at the same rate as before but would
gain mass at a lower rate.
5.
Illustrative Scenarios
We now utilize the framework developed above to construct models that illustrate different
plausible scenarios for the evolution of the quasar population. The simplest scenario is that the
luminous quasar population is dominated by black holes accreting with thin-disk efficiency ǫ0.1 ≈ 1,
all radiating with a "standard" SED (e.g., Elvis et al. 1994), and that the growth of black holes is
driven by the observed accretion. The key parameter of this scenario is the typical quasar lifetime,
which is linked in turn to the space density of black holes. However, there are many potential
variations on this theme, including the possibility that a large fraction of quasar activity is obscured,
that black hole growth is substantially affected by low redshift mergers, or that substantial black
hole growth occurs though low efficiency ADAF accretion. Our models here illustrate each of these
physically distinct possibilities, including long and short quasar lifetimes for the simplest scenario.
For simplicity, we will consider only models with p( mz) independent of mass. We examine only the
regime from z = 2 to z = 0 and choose parameters so that each model approximately reproduces
the observed optical QLF. Our goal is to determine what other observables, such as the QLF in
other bands, the masses of active black holes at different luminosities, the black hole mass function
itself, and the space density of host systems are most likely to discriminate among these scenarios.
We include some comparisons to recent estimates of these observables, but our emphasis is mainly
on the differences among the models themselves, since we have not made any adjustments to model
parameters to try to match these other data.
5.1. Model Parameters
The parameters of the five models are summarized in Table 2. Our baseline parameters are
similar to those used in the evolutionary calculations of §4.2.1. We adopt a double power-law n(M )
with slopes α = −1.5 and β = −3.4, which are required to match the asymptotic slopes of the Boyle
-- 39 --
et al. (2000) QLF for cases where p( m) is independent of mass. We adopt M∗ ≈ 109M⊙, making
the Eddington luminosity lM∗ close to the observed break luminosity at z = 2. Except for the
ADAF model (discussed below), we adopt a double power-law p( m) with parameters a = −0.5 and
b = −3.0. We start with a characteristic accretion rate m∗ = 0.5 and evolve it to lower redshift as
m∗ ∝ tγm. The normalization p∗ evolves as tγp, with different γp values from z = 2 to z = 1, z = 1
to 0.5, and z = 0.5 to 0. The values of γm for each model are chosen to give a reasonable match to
the observed evolution of Lbrk given the model's predicted growth of n(M ), and the values of γp are
then chosen by matching the observed amplitude of the QLF. Table 3 summarizes the quantitative
evolution of the five models, giving the values of M∗, the comoving black hole mass density ρbh,
and the mean accretion rate h mi at redshifts 2, 1, 0.5, and 0, and the accretion weighted lifetime
2 h m(t)idt at z = 1, 0.5, and 0.
tacc =R z
For our short quasar lifetime (short-tq) model, we normalize the black hole mass function at
z = 2 to n∗M∗ = 1.2 × 10−4 Mpc−3. Because of the high space density, the normalization of
p( m) is relatively low, and the accretion weighted lifetime between z = 2 and z = 0 is tacc =
9.5 × 106 yr, which implies little growth of black hole masses over this redshift range (see §4.1).
For the long-tq model, we reduce the black hole space density at z = 2 by a factor of ∼ 6, to
n∗M∗ = 2.012 × 10−5 Mpc−3, thus continuing the growing phase model illustrated in the lower
panels of Figure 7. The corresponding value of tacc is 5.3 × 107 yr, implying about one e-fold of
mass growth from z = 2 to z = 0 (compared to an order of magnitude growth from z = 5 to z = 2,
as shown in Figure 7). The factor of six difference in lifetimes is small compared to the full range
of quasar lifetimes discussed in the recent literature (Martini 2003 and references therein), but the
two models straddle the boundary between significant low-z accretion growth and minimal low-z
accretion growth. Solid and dotted curves in Figure 8 show the B-band QLFs predicted by the
short- and long-tq models, respectively. By construction, they match the Boyle et al. (2000) data
well at z = 2, 1, and 0.5.
The z = 0 data in Figure 8 come from Wisotzki (2000), based on the Hamburg/ESO quasar
survey. They lie close to an extrapolation of the Boyle et al. (2000) evolution model to z = 0. We
have chosen γp values so that each model passes through these data at MB ≈ −22, but we have not
attempted to reproduce the shape, so the model predictions do not nearly overlap as they do at
higher redshift. The short-tq model agrees well with the Wisotzki (2000) data. The long-tq model
predicts a higher amplitude of the QLF at high luminosities, mainly because greater black hole
growth gives it a higher value of M∗ at z = 0, and its luminosity function is steeper than the data.
The discrepancy could be reduced if we allowed m∗ to drop more rapidly between z = 0.5 and 0,
instead of extrapolating the behavior that fits from z = 2 to z = 0.5.
For the obscured model, we essentially take the short-tq model and multiply p∗ by five, assigning
20% of the active systems at each redshift the standard Elvis et al. (1994) SED and the remaining
80% the obscured SED of Table 1. The obscured model's optical luminosity function is shown
by the dashed lines in Figure 8, and it also closely matches the observed evolution. A steepening
of the evolution of p( m) to lower accretion rates is required to balance the increased black hole
-- 40 --
Table 2.
Input Model Parameters
M∗(z = 2)
n∗M∗(z = 2)
m∗(z = 2)
p∗(z = 2)
γm(z = 2 − 0)
γp(z = 2 − 1)
γp(z = 1 − 0.5)
γp(z = 0.5 − 0)
Short-tq
1
1.2 × 10−4
0.5
0.016
−2.7
2.84
2.72
3.75
Long-tq
1
2.012 × 10−5
0.5
0.097
−3.9
4.51
5.74
8.11
Obscured
1
1.2 × 10−4
0.5
0.08
−3.2
3.07
3.67
5.32
Merger
1
1.2 × 10−4
0.5
0.016
−3.9
5.33
5.71
8.83
ADAF
1.25
3.0 × 10−5
0.5
0.043
−3.1
1.60
2.79
3.50
Note. -- Defining parameters of the five illustrative models discussed in §5. All
models assume a double power-law n(M ) with α = −1.5, β = −3.4, and a mass-
independent, double power-law p( m) with a = −0.5, b = −3, except that the ADAF
model has a = −1.5 and a factor of 16 boost to p( m) in the range 10−4 ≤ m ≤ 10−2.
Models start at z = 2 with the tabulated values of M∗ (in 109M⊙), n∗M∗ (in comoving
Mpc−3 for Ωm = 1, h = 0.5),
m∗
evolves as tγm and p∗ as tγp. The obscured model has a 4:1 ratio of obscured to
unobscured systems. The merger model incorporates a factor fm = 1.8 growth by
equal mass mergers in each redshift interval z = 2 − 1, 1 − 0.5, and 0.5 − 0. Note that
in all five models the amplitude of p( m) drops with time at all
m even though the γp
values are positive, since m∗ decreases rapidly with time.
m∗, and p∗. Over the indicated redshift intervals,
-- 41 --
Table 3. Evolutionary Values
Short-tq Long-tq
Obscured
Merger ADAF
1
1.14
1.19
1.24
3.26
3.72
3.88
4.02
1
1.95
2.36
3.28
0.55
1.07
1.29
1.79
1
1.73
1.98
2.25
3.26
5.63
6.45
7.34
1
2.06
3.92
7.92
3.26
3.74
3.94
4.42
1.25
4.23
5.65
7.40
1.02
3.44
4.61
6.03
0.0054
0.0014
4.4 × 10−4
1.6 × 10−4
0.032
0.0054
0.0022
0.0027
0.027
0.0043
0.0014
7.1 × 10−4
0.0054
0.0015
6.1 × 10−4
0.0011
0.071
0.0084
0.0033
0.0015
0.59
0.79
0.95
3.01
3.86
5.35
2.46
3.08
3.66
0.62
0.86
1.38
5.45
6.79
8.00
M∗/109M⊙
z = 2
z = 1
z = 0.5
z = 0
ρbh/105M⊙Mpc−3
z = 2
z = 1
z = 0.5
z = 0
h mi
z = 2
z = 1
z = 0.5
z = 0
tacc/107yr
z = 2 − 1
z = 2 − 0.5
z = 2 − 0
-- 42 --
Fig. 8. -- Evolution of the B-band QLF for the five illustrative models discussed in §5. In each
panel, solid and dotted lines show results for two models dominated by thin-disk accretion, with
short and long quasar lifetimes, respectively. Short-dashed lines show a model with a large fraction
of obscured quasars, long-dashed lines a model in which mergers contribute substantially to the
evolution of the black hole mass function, and dot-dashed lines a model with high probability of
low m accretion, leading to significant black hole growth in an ADAF mode. Open circles in the
z = 2, 1, and 0.5 panels show the Boyle et al. (2000) QLF fit over the observed range of luminosities
at the indicated redshift. Asterisks in the z = 0 panel show the QLF estimate of Wisotzki (2000).
-- 43 --
growth from obscured accretion. Note that we could also have implemented the obscured scenario
by increasing the n(M ) normalization n∗ by a factor of five at z = 2 and keeping p∗ the same, but
then ρbh(z = 0) would have been very high.
For the merger model, we take the same initial parameters as the short-tq model and use
equation (37) to calculate black hole growth, with a merging factor fm = 1.8 between each pair
of redshifts shown in Figure 8 (2 → 1, 1 → 0.5, 0.5 → 0). This simple prescription for mergers is
not fully self-consistent because the luminosity function necessarily implies some accretion growth
as well. At each redshift, therefore, we calculate the mean accretion rate adopted to produce
the observed QLF and shift the black hole mass function by the corresponding amount. With this
additional accretion growth, M∗ increases by a factor of two over each of the three redshift intervals,
and by a factor of eight over the full range z = 2−0. The merger model's optical luminosity function
is shown by long dashed lines in Figure 8. Matching the observations requires steep evolution of
p( m) to compensate for the large amount of black hole growth from mergers. The doubling of M∗
between z = 0.5 and z = 0 leaves the merger model with a high amplitude tail of luminous systems
at z = 0.
The goal of our ADAF model, shown by the dot-dashed lines in Figure 8, is to illustrate a case
in which black holes experience substantial growth through low efficiency accretion at low redshift.
This requires a high probability of having m < mcrit, which is difficult to achieve while staying
consistent with the observed QLF. In particular, we are unable to find an acceptable fit to the
optical QLF by simply adjusting the parameters m∗ and a of our usual double power-law p( m).
After some experimentation, we settled on a model with the combination of a steeper low- m slope,
a = −1.5, and a boost to the probability of accretion rates below mcrit by a factor of sixteen. We
slightly increased M∗ to 1.25 × 109M⊙ to improve the match to the QLF break given our changed
p( m), and we reduced n∗M∗ to 3×10−5 Mpc−3 so that there would be more overall growth of n(M ).
With these choices, ρbh grows by a factor of about six from z = 2 to z = 0, and roughly two-thirds
of this growth comes from objects in an ADAF mode. The optical QLF is still significantly different
from that of the other models, but mostly at luminosities below the observed range. At z = 2,
the increased number density of low luminosity objects is mainly due to the steep slope of p( m),
which produces many faint thin-disk systems, but at lower redshifts the ADAF mode is directly
responsible for this excess of faint systems. The fact that we had to adopt such an artificial p( m) to
obtain a model that is even approximately consistent with the optical QLF already suggests that
low-z ADAF growth of black hole masses is not important in the real universe, but it is interesting
to explore the predictions of such a model nonetheless.
5.2. Black hole mass functions
We expect the black hole mass function n(M, z) to be a good discriminant among our models
because they involve different amounts of black hole growth, and in some cases start from different
n(M ) at z = 2. The best prospects for measurements of n(M ) are at z = 0, using the observed
-- 44 --
distribution of bulge luminosities or velocity dispersions and the observed correlation of dynamical
black hole masses with these properties. A number of authors have estimated the black hole mass
function in this way (e.g., Salucci et al. 1999; Merritt & Ferrarese 2001; Yu & Tremaine 2002; Aller
& Richstone 2002), and improving determinations of the form and scatter of the Mbh − σ relation
(Gebhardt et al. 2000; Merritt & Ferrarese 2000) and of the distribution of bulge dispersions (Sheth
et al. 2003) should yield more accurate estimates in the near future. Recent estimates of the average
black hole mass density at z = 0 are ρbh = 2 − 3 × 105(h/0.7)2 M⊙ Mpc−3 (Yu & Tremaine 2002;
Aller & Richstone 2002). High redshift estimates of n(M ) will be difficult, since achievable angular
resolution is not sufficient to measure dynamical masses of quiescent black holes. The distribution
of masses of active systems at a given luminosity can be measured using reverberation mapping
or emission line widths, and we discuss its diagnostic power in §5.4 below. It may be possible to
estimate the underlying n(M ) by establishing a correlation between Mbh and host galaxy properties
using active systems, then proceeding as at low redshift, but the observational uncertainties are
likely to remain considerable.
We show the model black hole mass functions in Figure 9, but to make differences more visible
we divide through by the prediction of the short-tq model and plot the logarithm of the ratio,
∆ log n(M ) = log n(M ) − [log n(M )]short−tq . The short-tq model has little growth of the black hole
mass function, with a change from M∗ = 1 × 109M⊙ at z = 2 to M∗ = 1.17 × 109M⊙ at z = 0. The
long-tq model at z = 2 has n(M ) a factor of ∼ 6 lower at all masses because it has the same value
of M∗ and a lower value of n∗. There is more growth in the long-tq model due to a higher p( m),
and the values of n(M ) begin to catch up to the short-tq case. Since the growth has the effect of
shifting the break in the double power-law n(M ) to higher masses, the change in n(M ) is larger at
the steep, high mass end and smaller at low masses, producing the characteristic kinked shape of
the lines in Figure 9. By z = 0 the long-tq model has overtaken the short-tq model at high masses,
but it remains below at low masses.
The final black hole mass densities for the short-tq and long-tq models are ρbh =R ∞
0 M n(M )dM =
is expected from the Soltan (1982) argument, which implies that ∆ρbh = R z2
4.02 × 105M⊙Mpc−3 and 1.80 × 105M⊙Mpc−3, respectively. The mass density added between z = 2
and z = 0.5 is nearly the same in the two models, ∆ρbh ≈ 0.7 × 105M⊙Mpc−3. This agreement
U (t)/(ǫc2)dt, where
U (t) is the bolometric emissivity of the quasar population at time t. Since mass growth in both
models is dominated by thin-disk systems with the Elvis et al. (1994) SED and ǫ0.1 = 0.1, and
both models reproduce the observed optical QLF, they necessarily have similar emissivity histories
and mean efficiencies and therefore similar ∆ρbh. The long-tq model adds significantly more mass
between z = 0.5 and z = 0, partly because it has a higher optical QLF at low redshift (Fig. 8),
and partly because its m∗ falls below mcrit, increasing the amount of low efficiency accretion. The
short-tq model has a final ρbh that is high in comparison with recent estimates, though arguably in
the range of their uncertainties. The long-tq model's ρbh agrees well with these estimates, coming
in slightly on the low side.
z1
The obscured model starts at z = 2 with the same n(M ) as the short-tq model, but n(M )
-- 45 --
Fig. 9. -- Evolution of the black hole mass function for the five models discussed in §5. Lines in
each panel show ∆ log n ≡ log10[n(M )/ns(M )], where ns(M ) is the mass function of the short-tq
model at the indicated redshift and n(M ) is the mass function of the long-tq, obscured, merger, or
ADAF model (dotted, short-dashed, long-dashed, and dot-dashed, respectively). The z = 0 panel
also lists the final value of the black hole mass density for each model in units of 105M⊙ Mpc−3. All
models have a double power-law mass function with slopes α = −1.5 and β = −3.4 at all redshifts,
and values of M∗(z) are listed in Table 3.
-- 46 --
grows quickly because of the large amount of optically invisible accretion. The break mass M∗
and mass density ρbh grow by a factor of 2.25 between z = 2 and z = 0, with a final ρbh about a
factor of two larger than the short-tq case and outside the range of recent estimates. The amount
of mass added, ∆ρbh ≈ 4 × 105M⊙Mpc−3, is about five times the amount for the short-tq model,
which is as expected because they have similar optical QLFs while the obscured model has four
optically invisible systems for each unobscured system. Since the observed optical QLF provides a
fairly natural fit to the estimate ρbh on its own (e.g., Yu & Tremaine 2002), it is difficult to add a
large amount of hidden accretion without overrunning these estimates. Higher efficiency accretion,
perhaps from spinning black holes, is one option (Elvis et al. 2002). However, part of the solution is
probably that our assumption of an 80% obscured fraction at all redshifts and luminosities, taken
from Comastri et al. (1995) and Fabian & Iwasawa (1999), is too extreme. Recent studies show
that faint X-ray sources are generally at lower redshifts than synthesis models predict, in which case
a smaller fraction of obscured sources are needed to produce the hard X-ray background (Barger
et al. 2002; Ueda et al. 2003).
The merger model starts with the same black hole mass function as the short-tq case, but it
evolves primarily by merging two lower mass black holes that create one higher mass black hole.
By z = 0, high mass black holes are more numerous than in the short-tq case by a factor of ten,
and low mass black holes are less numerous by a factor of three. The accretion growth in this
model is still dominated by thin-disk systems, and since mergers do not add mass to the black hole
population, the growth of ρbh tracks that of the short-tq and long-tq models down to z = 0.5. Like
the long-tq model, the merger model adds more mass at z < 0.5 because of its high QLF and low
m∗.
The ADAF model starts with M∗ = 1.25 × 109 M⊙ and a normalization of n(M ) a factor
of four lower than in the short-tq case. The high amount of accretion at low m values, in both
the thin-disk and ADAF regimes, results in more black hole growth than in any of the other pure
accretion models. By z = 1, n(M ) is similar to that of the short-tq case at low masses, and it is
a factor of ten larger at high masses. (Recall that all masses grow by the same factor even in the
ADAF model, so this difference just reflects the slope of the mass function in the two regimes.)
The final black hole mass density is 50% larger than that of the short-tq model, and it is difficult
to make parameter changes that significantly reduce this value because of the emissivity argument;
the optical QLF of this model traces the observations at observed luminosities, but it has a high
amplitude tail at low luminosity, and the mean efficiency is low because of the high fraction of
ADAF accretion.
Figure 10 compares the z = 0 black hole mass functions of the five models to an estimate
derived by combining the Sheth et al. (2003) estimate of the distribution of early-type galaxy
velocity dispersions with the M − σ relation found by Tremaine et al. (2002), log Mbh/M⊙ =
8.13 + 4.02 log(σ/200 km s−1). Filled triangles show the case where there is no intrinsic scatter
in the M − σ relation, while filled circles and squares show results assuming a log-normal p(M σ)
distribution with intrinsic scatter of 0.25 dex and 0.5 dex, respectively. Above 109M⊙, the derived
-- 47 --
Fig. 10. -- The black hole mass function at z = 0. Lines represent the five model mass mass
functions. Solid points show the mass function derived by combining the Sheth et al. (2003)
velocity dispersion distribution of early-type galaxies with Tremaine et al.'s (2002) estimate of the
M − σ relation between black hole mass and bulge velocity dispersion. Triangles, circles, and
squares show results assuming no intrinsic scatter in the M − σ relation, 0.25-dex scatter, and
0.5-dex scatter, respectively. Open circles show the effect of adding Aller & Richstone's (2002)
estimated contribution from spiral galaxies to the 0.25-dex scatter (filled circle) estimate for early-
type galaxies.
-- 48 --
mass function is quite sensitive to the assumed intrinsic scatter. Tremaine et al. (2002) argue
that this scatter is no larger than ∼ 0.25 − 0.3 dex, but the data are concentrated in the range
107 − 108.5M⊙, and the scatter could increase or decrease with mass. Above 109.5M⊙, the derived
mass function also depends on extrapolating the mean M − σ relation beyond the range of current
observations. Since Sheth et al. (2003) consider only early-type galaxies, we show with open circles
the effect of adding Aller & Richstone's (2002) estimate of the spiral galaxy contribution, which
becomes important below ∼ 107.8M⊙. If we included their S0 contribution as well (which might
double count galaxies already in the Sheth et al. sample), then the mass function would be higher
by 0.2 dex in this regime.
If the extrapolation of the M − σ relation is correct and the scatter is indeed . 0.3 dex, then
even the short-tq model overpredicts n(M ) above 109M⊙, and our other models fare worse because
of their higher values of M∗(z = 0). Bringing the models in line with this estimate of n(M ) would
require either reducing our initial M∗(z = 2) by 0.5-1 dex or changing our double power-law form
of n(M ). We selected M∗(z = 2) = 109M⊙ because the Eddington luminosity lM∗ is then close to
Boyle et al.'s (2000) break luminosity (more precisely, lM∗ = 1.7Lbrk), making it straightforward to
fit the observed Φ(L). However, given the interplay between n(M ) and p( m), there is at least some
room to reduce M∗ and continue to match the observed QLF without requiring super-Eddington
luminosities. If we instead adopted an exponential high-M cutoff (or a steeper high-M power-law
slope), then we would need a mass-dependent p( m) to reproduce the Boyle et al. (2000) data at
z = 2, with more massive black holes having a higher probability of being active. We will explore
these implications, and the tradeoff with uncertainties in the data, in future work, where we run
models backwards from the z = 0 mass function instead of forwards from the z = 2 QLF.
5.3. X-ray and FIR luminosity functions
In three of our models (short-tq, long-tq, and merger), the bolometric emission of the quasar
population is dominated by systems with a thin-disk SED. We therefore expect them to make sim-
ilar predictions for the QLF in all wavebands (since they match in the B-band by construction).
However, the obscured and ADAF models have large populations of systems with different SED
shapes, and they may be distinguished by their X-ray or FIR luminosity functions. Observation-
ally, the X-ray luminosity function is not as well characterized as the optical luminosity function,
especially at high redshifts, but large surveys following up sources from ROSAT, ASCA, Chandra,
and XMM-Newton are transforming the situation and yielding much better constraints on the evo-
lution of the X-ray QLF (e.g., Miyaji, Hasinger, & Schmidt 2001; Cowie et al. 2003; Fiore et al.
2003; Hasinger 2003; Steffen et al. 2003; Ueda et al. 2003). In the mid/far-IR, SCUBA detects
the brightest sources at 850µm (Priddey et al. 2003), but the revolutionary instrument should
be SIRTF, with a much greater combination of wavelength range and sensitivity than previously
available.
We calculate X-ray and FIR luminosity functions of our five models using the Fν values in
-- 49 --
Table 1 for systems with the various accretion modes. For X-ray QLFs we use observed-frame
bandpasses of 0.5 − 2 keV and 2 − 10 keV at redshifts z = 2, 1, and 0.5. Results are shown in
Figure 11. To enhance the visibility of model differences, we again plot the log of the ratio of each
model's predictions to those of the short-tq model. The short-tq model's predictions at z = 2 are
nearly identical to those shown by the solid lines in Figure 3. As expected, results for the long-tq
and merger models are very similar to short-tq in all bands because of the dominance of thin-disk
SEDs, and they could probably be brought closer still with slight adjustments of model parameters.
The low-z X-ray luminosity functions of the long-tq and merger models are slightly enhanced at
low luminosities relative to B-band because their low m∗ values (an indirect consequence of higher
M∗) lead to more ADAF accretion.
The obscured model, which initially has five times as many accreting systems as the short-tq
model, shows nearly this full factor of five enhancement in the 2 − 10 keV band at z = 2, where
obscuration is almost negligible. This enhancement shrinks steadily towards lower z, especially at
higher luminosities, as the observed-frame 2 − 10 keV band becomes more affected by obscuration
(see Table 1). The 0.5 − 2 keV band shows a significant (0.4-dex) enhancement at low luminosities
at z = 2, comprised of systems that have high bolometric luminosity and are thus able to shine
detectably at this wavelength despite obscuration. However, at lower redshift the 0.5 − 2 keV
band is almost completely extinguished, and the obscured QLF is no different from that of the
short-tq model. The most dramatic feature of the obscured model is the booming FIR luminosity
function, enhanced by 1 − 2 orders of magnitude at all redshifts because the numerous obscured
systems re-radiate all of their absorbed UV and soft X-ray luminosity in the FIR. This distinctive
prediction of models with a large obscured population should be easily testable with SIRTF. The
prediction holds regardless of whether obscured and unobscured systems represent two separate
populations or different orientations of the same population, provided that the absorbed energy is
indeed re-radiated.
As previously noted, the ADAF model has some substantial differences from the short-tq model
even in B-band. At z = 2, the steady rise in low luminosity systems is a consequence of the steeper
slope of p( m), while the bump at log L ≤ 1043.5erg sec−1 reflects the boosted number of objects with
m < mcrit and thus consists of objects accreting in an ADAF mode. At low redshifts, this ADAF
bump moves to slightly higher luminosities. The B-band differences reappear at other wavelengths,
but there are further differences in the X-ray bands that reflect the larger fraction of the ADAF
SED that emerges in these bands. At high redshift, the low luminosity boost to the QLF is larger
in X-ray than in optical, exceeding an order of magnitude. At low redshift, the decreasing value
of
m∗ leads to a still higher probability of ADAF accretion, and high mass black holes in ADAF
mode boost the X-ray QLF even at high luminosities. Indeed, we can infer from Figure 11 another
prediction of our ADAF model: at z . 1, most X-ray selected AGN should be ADAF systems,
at every luminosity. This does not appear to be the case in the real universe, providing a further
observational argument against the importance of low efficiency accretion as a black hole growth
mechanism at low redshift.
-- 50 --
Fig. 11. -- Ratios of the luminosity functions of the long-tq, obscured, merger, and ADAF models
to those of the short-tq model at z = 2, 1, and 0.5 (left, middle, right). Lines in each panel show
∆ log Φ ≡ log10[Φ(L)/Φs(L)] for B-band, 2-10 KeV, 0.5-2 KeV, and 10-1000 µm (top to bottom).
The X-ray bands are observed-frame with the redshift effects on Fν shown in Table 1 included.
-- 51 --
Fig. 12. -- The observed frame soft X-ray luminosity function at z =2,1,0.5, and 0. Lines show the
predictions of our five models, as indicated. Open circles show the evolutionary model fits to the
ROSAT 0.5-2 keV QLF from Miyaji, Hasinger, & Schmidt (2001), for Ωm = 1, h = 0.5. Points are
plotted over the range of luminosities spanned by the data at each redshift.
-- 52 --
Fig. 13. -- The intrinsic hard X-ray luminosity function at z =2, 1, 0.5, and 0. Lines show the
predictions of our five models, as indicated. Open circles show the evolutionary model fits to the
2-10 keV QLF from Ueda et al. (2003), based on HEAO1, ASCA, and Chandra surveys. Ueda
et al. (2003) use spectral modeling to estimate the rest-frame 2-10 keV luminosity corrected for
obscuration. We therefore use the 2-10 keV Fν values from Table 1 at each redshift, and we use
the thin-disk values of Fν for obscured systems in the obscured model, since these represent the
intrinsic luminosities.
-- 53 --
Figures 12 and 13 compare the model predictions to estimates of the soft and hard X-ray
luminosity functions from Miyaji, Hasinger, & Schmidt (2001) and Ueda et al. (2003), respectively.
As with the optical QLF, we plot the authors' evolutionary model fits over approximately the range
covered by the observational data at each redshift. Miyaji et al.'s (2001) luminosity function, is
based on observed-frame, 0.5-2 keV luminosities from the ROSAT All Sky Survey, with no correction
for X-ray obscuration. Ueda et al. (2003), on the other hand, use spectral shape information to
estimate the intrinsic (i.e., corrected for obscuration), rest-frame 2-10 keV luminosity function, from
a combination of HEAO1, ASCA, and Chandra surveys. We compute the corresponding quantities
from our models in each case.
At z = 2, all models fit the high luminosity end of the 2-10 keV QLF, except for the obscured
model, which overpredicts by a factor of five. However, all models overpredict the 0.5-2 keV QLF
by a factor ∼ 3. At z = 1 and z = 0.5, models come into better agreement with the 0.5-2 keV QLF
(except for the ADAF model, which remains high), but the thin-disk dominated models (short-tq,
long-tq, merger) fall below at 2-10 keV. By z = 0, the M∗ growth in the long-tq and merger models
has brought them back into better agreement at 2-10 keV, but they are high at 0.5-2 keV, while
the short-tq model is about right at 0.5-2 keV and well below at 2-10 keV. The obscured model is
in rough agreement with both X-ray QLFs at z ≤ 1.
In brief, the situation is confusing, and there is no single obvious change that would bring
any of the models into agreement with both X-ray luminosity functions and the B-band luminosity
function (Fig. 8) at all redshifts. A more sophisticated obscuration model, with soft X-ray and
optical obscuration becoming more important at low redshifts and low luminosities, would certainly
help, and it might explain why the faint end of the 2-10 keV QLF rises above the B-band QLF
at z = 2. However, the overprediction of the soft X-ray QLF at z = 2 by models that match
the B-band and 2-10 keV QLF seems difficult to understand.
It is worth noting that Ueda et
al.'s models, which incorporate luminosity-dependent obscuration, also tend to overpredict the soft
X-ray QLF (see their figure 15), and that Miyaji et al.'s redshift bins become large at high redshift
(e.g., z = 1.6 − 2.3 and 2.3 − 4.6), which may make the model interpolation to a given redshift less
accurate.
5.4. Masses and accretion rates of active black holes
While the underlying black hole mass function n(M ) may be difficult to determine at z > 0,
the distribution of active black hole masses is more accessible. As discussed in §3.3, the mass
distribution of active systems depends on both n(M ) and p( m), and its variation with luminosity
can be a valuable discriminant of models. Locally, the masses of active systems can be measured
by combining emission line widths with sizes of the emitting regions estimated by reverberation
mapping (Wandel et al. 1999). This approach can in principle be extended to high redshift, but
fainter targets and longer variability timescales make it difficult. A more broadly applicable method
is to combine line widths (e.g., Hβ, or C IV at higher redshift) with the average size-luminosity
-- 54 --
relation inferred from reverberation mapping of local objects or from photoionization modeling (e.g.,
Laor 1998; Gebhardt et al. 2000; McLure & Dunlop 2002; Corbett et al. 2003; Vestergaard 2004).
In the last few years, these methods have yielded black hole mass estimates over an increasing
range of redshift and luminosity (e.g., Woo & Urry 2002). Estimates of mass accretion rates are
typically made by combining mass estimates with luminosities for an assumed efficiency, so they
are not independent of the mass estimates themselves. However, quasar SEDs may also provide
at least rough diagnostics of accretion rates, in the broad categorization of thin-disk vs. ADAF
systems, for example, and perhaps in the finer distinction between near-Eddington systems and
significantly sub-Eddington systems (Kuraszkiewicz et al. 2000; Czerny et al. 2003).
Figure 14 illustrates the range of black hole masses that contribute to different ranges of B-
band luminosity at z = 2, 1, 0.5, and 0. For each redshift and model, a symbol marks the median
mass Mmed of black holes that are active in this luminosity range, and a vertical bar shows the
10% -- 90% range of masses at this luminosity. Small black dots show the model's value of M∗, the
mass of the break in the black hole mass function. Different panels represent different luminosity
ranges, which we express in terms of Lbrk, the break parameter in the observed B-band QLF. Bear
in mind that the physical luminosity associated with Lbrk decreases towards low redshift, and that
Lbrk drops increasingly below lM∗. Furthermore, the z = 0 results here and in subsequent plots
should be taken with a grain of salt because we do not require our models to match the shape of
the QLF at z = 0, only the normalization at Lbrk.
Figure 14 is interesting both for the features that are common to all of the models and for the
features that distinguish them. The key common feature is a change in the relation between mass
and luminosity from high redshift to low redshift. At z = 2, in all models, the sequence of luminosity
is also a sequence of black hole mass: the median active mass rises from Mmed ≈ 5.6× 108M⊙ in the
lowest luminosity range to Mmed = 6.8 × 109M⊙ in the highest luminosity range, roughly the same
factor by which the luminosity itself rises. Comparing the symbols to the black dots shows that low
luminosity quasars arise from sub-M∗ black holes and high luminosity quasars from super-M∗ black
holes. The 10% -- 90% range of masses in a given luminosity range is only about 0.4-dex, similar
to the width of the luminosity bin itself. At low redshift, on the other hand, the trend of median
mass with luminosity is much weaker, and even L > 6.25Lbrk systems have median masses lower
than M∗. The distribution of masses for a given luminosity range is much broader than at high
redshift, typically close to an order of magnitude. As we show more directly in Figure 15 below,
the luminosity function at low redshift represents largely a sequence of accretion rate rather than
black hole mass.
The differences between models largely trace the differences in the growth of M∗. At z = 2,
Mmed is similar at a given luminosity for all models. At lower redshifts, the short-tq model, which
has the least M∗ growth, always has the lowest Mmed at a given luminosity, followed by the obscured
model, which has the next lowest growth of M∗. The merger and ADAF models have the most M∗
growth and the highest Mmed values at low redshift. The ADAF model is generally highest, even
though the merger model overtakes it in M∗ by z = 0, because the high probability of low m favors
-- 55 --
Fig. 14. -- Evolution of the mass distribution of active black holes as a function of luminosity for the
five illustrative models discussed in §5. Each panel corresponds to the range in B-band luminosity
shown in the top center relative to the QLF break luminosity Lbrk(z). The small black dots show
the values of M∗(z) for each model, and the open symbols show the median mass of black holes
contributing to the given luminosity range for the short-tq, long-tq, obscured, merger, and ADAF
models (triangle, square, pentagon, circle, and star). The vertical bars show the 10%-90% range
of the distribution. The horizontal offset of the points from z = 2, 1, 0.5, and 0 is artificial and is
done to distinguish the models from each other.
-- 56 --
high mass black holes at a given luminosity. The 10% -- 90% spread at a given redshift is generally
similar among the models.
The trends and model differences that we have shown here for B-band luminosity generally
hold for luminosities defined in other bands as well. The most significant change is that the median
black hole mass at fixed (in erg sec−1) FIR luminosity is much smaller in the obscured model than
in all other models because a large fraction of the obscured SED emerges in the FIR band, enabling
low mass black holes to produce high luminosity. The other significant change is that the median
masses for the ADAF model are considerably lower in both X-ray bands because ADAF accretors
emit a larger fraction of their bolometric energy in X-rays.
While the distributions of black hole masses and of accretion rates associated with a given
luminosity provide essentially the same information, it is helpful to look directly at both distribu-
tions. Figure 15 is similar in spirit to Figure 14, but it shows the distribution of accretion rates at a
given luminosity rather than the distribution of black hole masses. Symbols mark the median value
of
m, vertical bars show the 10% -- 90% range of the distribution, and small black dots show each
model's m∗ at each redshift. A model-to-model comparison shows essentially the reverse behavior
from Figure 14, with higher median black hole masses corresponding to lower median accretion
rates. The key results are that the median m is close to m∗ at all luminosities at z = 2, in all of the
models, while at low redshift the median m is an increasing function of luminosity. Furthermore,
even the low luminosity systems tend to have m > m∗ at low redshift.
The ADAF model is the outlier in this plot because its p( m) distribution is strongly skewed to
favor low accretion rates. It consistently has the lowest median accretion rate at a given luminosity
and redshift, and the spread in accretion rates is large. At z = 0.5, the median m in the 0.4Lbrk −
Lbrk luminosity bin is equal to the critical value mcrit at which ADAF accretion sets in, indicating
that about half of optically systems selected in this luminosity range are predicted to be ADAF
accretors. A similar conclusion holds even for the highest luminosity bin at z = 0. The situation is
more extreme for X-ray selection, where the ADAF model predicts a median m in the ADAF range
in all luminosity ranges at z ≤ 1. As already noted in §5.3, this prediction that a large fraction of
luminous X-ray quasars are ADAF systems appears to be an empirical failure of this model.
All five of our models assume p( m) independent of mass, and the transition in behavior from
m∗ that is required
high redshift to low redshift is a consequence of the declining evolution of
to match the Boyle et al. (2000) luminosity evolution in any such model. However, as shown in
Figure 6, it is also possible to reproduce the Boyle et al. (2000) results with a model in which
m∗ is constant but p( mM, z) is mass-dependent. Since this model is dominated by thin-disk
accretors, its luminosity function is similar to that of our short-tq model in every band. However,
Figure 16 shows that its predicted distribution of active black hole masses is strikingly different
at low redshift. Because the mass-dependent model reproduces the downward evolution of Lbrk
by preferentially reducing activity in massive systems, it predicts much lower median black hole
masses at any luminosity at low redshift. Furthermore, the luminosity function remains primarily
-- 57 --
Fig. 15. -- Evolution of the accretion rate distribution of active black holes as a function of lu-
minosity for the five illustrative models discussed in §5. This figure is similar to Figure 14, but
presents results in terms of accretion rates rather than black hole masses. Each panel corresponds
to the range in B-band luminosity shown in the top center. The small black dots show the value
m∗(z) for each model, open symbols show the median accretion rate contributing to the given
of
luminosity range, and the vertical bars show the 10%-90% range of
m. The offset of the points
from z = 2, 1, 0.5, and 0 is artificial and is done to distinguish the models from each other. The
values of m∗ at z = 0 for the long-tq and merger model fall well below 10−3 and are off the bottom
of the plot.
-- 58 --
a sequence of black hole mass even at low redshift, with the median mass rising by a factor of ten
between our lowest and highest luminosity bins at z = 0, compared to only a factor of three for the
short-tq model. Because of the tight link between black hole mass and luminosity in this model,
the spread in masses at a given luminosity remains small even at low redshift. The evolution of
the distribution of active black hole masses thus provides an excellent tool for deciding whether the
observed decline of Lbrk reflects decreasing characteristic accretion rates or a preferential drop in
activity among more massive black holes.
Recent studies using line widths to estimate black hole masses for large data samples (e.g.,
McLure & Dunlop 2003; Vestergaard 2004) provide the kind of data needed to test the predictions
in Figure 14 -- 16. Vestergaard's (2004) Figure 5a bins estimated black hole masses by luminosity and
redshift, allowing a qualitative comparison. At z ∼ 2, the typical estimated masses for L ∼ Lbrk are
∼ 109M⊙, in agreement with our model initial conditions, and there is a clear trend of estimated
black hole mass with luminosity, though perhaps less strong than the nearly linear relation predicted
by our models. At z ∼ 0, there is a broader spread in estimated masses at a given luminosity, as
predicted by our standard models in which m∗ declines at low z. However, the typical mass at
L ∼ Lbrk is ∼ 107.5 − 108M⊙, which is in between the predictions of the mass-independent p( m)
and mass-dependent p( m) models shown in Figure 16. Careful assessment and modeling of the
statistical errors is needed to draw reliable conclusions from a more quantitative comparison, since
the random errors in the mass estimates are large enough (a factor ∼ 3) to distort the underlying
mass distributions significantly.
5.5. Space densities of quasar hosts
Some of our models have a low space density of black holes and a high duty cycle -- i.e.,
low n(M ) and high p( m) -- while others have more numerous black holes and lower duty cycles.
Unfortunately, neither the luminosity function nor the distribution of active black hole masses
distinguishes these cases, since both depend only on the product n(M )p( mM ) (see eqs. 8 and 21).
However, if the locally observed correlations between black hole mass and bulge velocity dispersion
or luminosity continue to higher redshifts, they offer a tool for diagnosing, at least approximately,
the underlying space density of black holes that shine at a given luminosity. If the space density is
low and the duty cycle high, then quasars should reside in rare host galaxies with luminous bulges.
If the space density is high, then host galaxies should include later type and less luminous systems.
To translate this idea into a precisely defined observable, we find the median mass of black
holes that produce quasars in a given luminosity range (the symbols in Figure 14), then compute
the space density of black holes with this mass or greater. The corresponding observational pro-
gram would require measuring the median host galaxy luminosity Lhost,med(Lq) of quasars in the
same luminosity range, then measuring the galaxy luminosity function at the same redshift and
computing Φ(L > Lhost,med), the space density of galaxies brighter than Lhost,med(Lq). The pre-
dicted and observable quantities are directly comparable if the scatter between Mbh and Lhost is
-- 59 --
Fig. 16. -- Evolution of the mass distribution of active black holes as a function of luminosity for
the short-tq model (triangles) and the mass-dependent p( mM ) model of §4.2.2 (squares). The
format is similar to that of Figure 14. Small black dots show values of M∗(z), which are nearly
identical for both models. Open points show the median mass of active black holes in the indicated
luminosity range, and vertical bars show the 10%-90% mass range.
-- 60 --
negligible. Note that this condition does not imply negligible scatter between Lq and Lhost, since
m variations still produce variations in quasar luminosity. Since black hole mass appears to be
most directly correlated with bulge properties, one would ideally use host bulge luminosity and
the bulge luminosity function rather than total luminosities. Accurate bulge-disk decomposition at
high redshift may be impractical, however, especially in the presence of an active nucleus, so the
next best thing is to use a red passband that is sensitive to old stellar populations. While the space
density of hosts is not as informative as an actual measurement of the black hole mass function
n(M, z), it is less demanding observationally, since it does not require calibration of the Mbh − Lhost
relation at high redshift, only the existence of a relation with relatively small scatter.
Figure 17 shows the model results for the luminosity ranges 0.4Lbrk < L < Lbrk and L >
6.25Lbrk at redshifts z = 2, 1, 0.5, and 0. As expected, the long-tq model starts with a host space
density that is a factor of ten below that of the short-tq model, and although it catches up at
lower redshift because of the greater amount of black hole growth, a considerable gap remains. The
obscured model starts with an n(M ) close to that of the short-tq model, and its predictions for
host space densities remain close to it at all redshifts, with the greater growth of black hole masses
largely compensated by the faster decline in m∗. Note, however, that this model's prediction would
have been quite different if we had implemented it by boosting the black hole space density instead
of the duty cycle.
The merger model is identical to the short-tq model at z = 2, but its distinctive evolution
of n(M ), with low mass black holes transforming into high mass black holes, leads to different
behavior of the predicted space densities with redshift and with luminosity.
In particular, the
low-z depletion of the low end of n(M ) leads to a relatively low space density, especially at low
luminosity. The ADAF model starts with a relatively low n(M ) (intermediate between long-tq and
short-tq) and thus a relatively low host space density. However, its mass function overtakes that
of short-tq at high masses by z = 1 (see Fig. 9), and the predicted host space densities are similar
for L > 6.25Lbrk. At low luminosities and higher redshifts, the ADAF model has a lower space
density despite its large amount of black hole growth because its p( m) distribution favors rarer,
higher mass black holes at a given luminosity.
The model that stands out most distinctively in Figure 17, at least in terms of its redshift
dependence, is the mass-dependent p( m) model, shown by the filled triangles. Because this model
matches the QLF by shifting activity preferentially towards low mass black holes at low redshift,
the predicted host space density climbs rapidly. In this model, the hosts of moderate luminosity
quasars at low redshift should include relatively late-type galaxies and low luminosity ellipticals.
In effect, the properties of hosts offer an indirect way to detect the sharp decline in typical active
black hole mass shown in Figure 16.
If the masses of black holes are correlated with the masses of the dark matter halos in which
they reside, a natural expectation given the observed correlation with bulge velocity dispersion, then
quasar clustering provides another observational tool for inferring their space density. A low black
-- 61 --
Fig. 17. -- The comoving space density of black holes with mass above the median mass of active
systems with B-band luminosity 0.4Lbrk < L < Lbrk (top) or L > 6.25Lbrk (bottom). Open
symbols show results for the five models discussed in §5, as indicated, and filled triangles show
results for the mass-dependent p( m) model illustrated in Figure 6. These predictions can be tested
by studies of quasar hosts or quasar clustering. Note the different y-axis ranges of the top and
bottom plots.
-- 62 --
hole space density implies that the host halos are rare, massive systems that tend to be strongly
clustered (Kaiser 1984; Mo & White 1996), while a high space density implies more common, less
strongly clustered hosts. This is the idea behind the proposals of Haiman & Hui (2001), Martini
& Weinberg (2001), and Kauffmann & Haehnelt (2002) to constrain quasar lifetimes (hence duty
cycles, hence black hole space densities) using the quasar correlation function or the quasar-galaxy
cross-correlation function (see also Haehnelt et al. 1998).
In Figure 17, models with low points
predict strong quasar clustering and models with high points predict weaker clustering. To a first
approximation, one could calculate the expected bias (relative to mass clustering) for quasars in the
two luminosity ranges by reading off the space density from the figure, finding the mass threshold
for halos that have this space density given an assumed cosmological model, and calculating the
bias of halos above this mass threshold using the methods of Mo & White (1996; see Martini &
Weinberg [2001] for a more detailed description of this approach). One could also carry out a more
thorough calculation, weighting the contribution to the bias by the fraction of black holes of a
given mass contributing to the luminosity range, but we suspect that the results would not be very
different.
Recent results from the 2dF Quasar Redshift Survey favor a relatively low comoving correlation
length, s0 ≈ 5.8h−1 Mpc in redshift space, with no clear evidence for dependence on redshift or
luminosity (Croom et al. 2003). Using the Martini & Weinberg (2001) model, which assumes simple
on-off quasar activity with a monotonic relation between quasar luminosity and host halo mass,
the implied lifetime is short, t ∼ 106 years. The current data set is not quite large enough to allow
a precise clustering measurement for a volume-limited subset of quasars at z ∼ 2 − 3, which is what
one would ideally like to use for the lifetime analysis. However, if future results continue to show a
low correlation length at high redshift, and no significant dependence on luminosity, then they may
indicate that there is substantial scatter in the relation between quasar luminosity and host halo
mass. This scatter could in turn indicate that the correlation between black hole mass and halo
mass at high redshift is much weaker than the measured correlation between black hole mass and
bulge mass at low redshift, or else that the luminous quasars have a wide range of L/LEdd even at
high-z. Clustering analyses of the full 2dF quasar survey and of the SDSS quasar survey should
yield interesting insights on these questions over the next few years.
6. Summary
In the framework developed here, the central actor in black hole and quasar evolution is
the accretion rate distribution p( mM, z), the probability that a black hole of mass M accretes
at a rate m (in Eddington units) at redshift z. Given a model for the accretion efficiency as a
function of
m, which can be inferred from observations and theoretical considerations that are
largely independent of QLF evolution per se, the combination of p( mM ) and the black hole mass
function n(M ) determines the bolometric luminosity function via equation (8). Furthermore, in
the absence of mergers, p( mM, z) determines the evolution of n(M, z) given a "boundary value"
-- 63 --
of n(M ) at some redshift. Mergers can complicate the picture by changing n(M ) independently
of p( mM, z). We have generally made the plausible but not incontrovertible assumption that
black holes accreting in the range 0.01 < m < 1 have "thin-disk" efficiencies ǫ0.1 ≈ 1 and that
efficiencies decrease at higher (super-Eddington) and lower (ADAF) accretion rates (eq. 7). We
have derived many of our results under the mathematically simplifying assumption that p( m) is
independent of mass. While this assumption is unlikely to hold to high accuracy, it may be a
reasonable approximation at redshifts near the peak of quasar activity, since black holes that grow
faster than their peers tend to reduce their accretion rates in Eddington units and vice versa. Most
of our specific examples assume that n(M ) and p( m) are double power-laws with breaks at M∗ and
m∗, respectively, with p( m) truncated at mmin = 10−4 and mmax = 10.
6.1. Basic Results
Our framework yields a number of mathematical results that give insight into the relations
among the black hole mass function, the accretion rate distribution, and the QLF. When p( m) is
independent of mass, the convolution integral (8) for Φ(L) can be understood as follows: for each
range m → m + d m, the mass function n(M ) is mapped to a luminosity L = ǫ0.1 mlM , multiplied
by p( m)d m, and added to a running total. If n(M ) is a double power-law and the range of
m is
bounded, then the asymptotic slopes of Φ(L) must equal the low and high mass slopes of n(M ),
since high luminosity objects must come from black holes with M > M∗ and low luminosity objects
must come from black holes with M < M∗. In the intermediate luminosity regime, where black
holes above and below M∗ can both contribute, the QLF turns over in a way that depends on the
slopes of n(M ) and the shape of p( m). For our usual double power-law p( m), the QLF break occurs
at a luminosity Lbrk ∼ m∗lM∗. If m∗ is close to one, then the break luminosity corresponds roughly
to the Eddington luminosity of M∗ objects, and the slope above the break corresponds to the high
mass slope of n(M ). However, if
m∗ is low, then the turnover may be associated largely with the
change in slope of p( m), and the asymptotic regime where the high-L slope of Φ(L) matches the
high-M slope of n(M ) may only be reached beyond the observed range of luminosities. For most
of the models that we have presented here, which are designed to match the Boyle et al. (2000)
optical luminosity function, the first case applies at high redshift and the second at low redshift.
A key feature of our model of accretion physics is that objects do not radiate at super-Eddington
luminosities -- instead, we assume that the efficiency is ǫ0.1 = m−1 for m > 1, so that super-
Eddington accretors radiate at L = LEdd. Although we have generally adopted m∗ = 0.5 − 1
for fitting data at z ≥ 2, our results would not be very different if we took m∗ > 1, or if we
changed the maximum accretion rate or the high- m slope, because the change in efficiency would
cut off the luminosity distribution at LEdd anyway. Our results would also not be very different
if we assumed that black holes fed at a super-Eddington rate by their host galaxy regulate their
accretion by outflows or convection (Blandford & Begelman 1999; Quataert & Gruzinov 2000) so
that they gain mass at m ≈ 1 and radiate at L ≈ LEdd. If such flows drove gas out into the galaxy
-- 64 --
halo, they would effectively truncate p( m) at m = 1, while if they returned gas to a reservoir from
which it would eventually be accreted, they would transform the m > 1 tail of p( m) into a spike at
m = 1. With our standard efficiency assumptions, the latter scenario would increase the average
radiative efficiency of high- m accretion, by moving it from the super-Eddington regime to the thin-
disk regime, thus yielding more luminosity for a given amount of black hole growth. Changing the
accretion physics to allow ǫ0.1 ≈ 1 with m > 1, and thus to allow substantially super-Eddington
luminosities when the accretion rate is high (Begelman 2002), would have a more drastic effect on
our results. In this case, only a truncation of p( m) would cut off the luminosity distribution at a
given black hole mass, and the predicted luminosity function would therefore be sensitive to the
values of
m∗ and mmax and to the shape of p( m) at m > 1.
The fraction of black holes of mass M that are active at a luminosity L is proportional to
L
n(M )p(cid:16) m =
ǫ0.1 mlM(cid:17), the underlying black hole space density times the probability of having the
accretion rate required to shine at L. The same product appears in the integrand for the luminosity
function itself (eq. 8), but by constraining the integrand rather than the integral, measurements of
active black hole masses can discriminate among models that produce similar Φ(L) with different
p( m) and n(M ). In terms of our double power-law models, the active mass distribution is particu-
larly useful for distinguishing cases with high M∗ and low m∗ from models with low M∗ and high
m∗.
A general mass-dependent p( m) can be written in the form p( mM ) = p0( m)D(M m), with
p0( m) ≡ p( mM0) for some fiducial mass M0 and D(M0 m) ≡ 1. With such a mass dependence,
we can still understand the convolution integral for Φ(L) by considering the contribution from each
range m → m + d m, but where the sum before involved only horizontally and vertically shifted
versions of the mass function n(M ), now the mass function can be tilted or distorted by D(M m)
before being added to the running total. Mass-dependence of p( m) thus breaks the tight link
between n(M ) and Φ(L) and adds freedom to models of the QLF. Specifically, for a model with
a given n(M ) and a mass-independent p( m), there is a family of models with different n(M ) and
mass-dependent p( m) that yield identical predictions for Φ(L) and the distribution of active black
hole masses. However, this degeneracy applies only at a single redshift; the evolution of models with
mass-dependent p( m) is different because the shape of n(M ) changes with time. Furthermore, while
the active black hole mass distributions are the same, the underlying black hole mass functions are
different, and a measurement of the full n(M ) at z = 0 may be sufficient to diagnose the mass-
dependence of p( m) at higher redshifts.
If two models have similar underlying n(M ), then the
masses of active black holes or space densities of their host galaxies are powerful diagnostics for
mass dependence of p( m), as illustrated in Figures 16 and 17.
For evolutionary calculations, a crucial simplification (eq. 10) is that the accretion driven
mp( mM, z)d m, not on
the full form of the distribution function. Between times t1 and t2, a black hole with accretion
mdt is the accretion weighted
lifetime (eq. 23) and ts is the Salpeter time (eq. 5). In any particular mass bin, some of the black
growth of n(M ) depends only on the mean accretion rate h m(M, z)i ≡R ∞
rate m(t) grows in mass by a factor exp(tacc/ts), where tacc = R t2
t1
0
-- 65 --
holes grow faster than average and some grow slower, but n(M ) evolves as if all of them grew at
the average rate for that bin. If h mi is independent of mass, then the growth factor is the same
for all mass bins, and the mass function simply shifts in a self-similar fashion, preserving its shape
(eq. 11). One might expect this self-similar behavior to emerge, approximately, as a "fixed point"
solution during the epoch of rapid black hole growth. If h mi depends on mass, then the shape of
n(M ) changes with time, becoming steeper if low mass black holes grow more rapidly and shallower
if high mass black holes grow more rapidly.
Our results also imply a number of critical values for the logarithmic slopes of p( m) and n(M ),
where the character of the solutions changes. We summarize these critical slopes in Appendix B.
The most important result of this sort is that growth by accretion drives n(M ) at fixed mass up
if the mass function is steeper than M −1 and down if it is shallower, and that the corresponding
critical slope for merger driven growth is −2 rather than −1. Generically, one expects mergers to
increase n(M ) at high masses and decrease it at low masses.
6.2.
Implications of the Observed Luminosity Evolution
We have kept the comparison to observations in this paper rather loose, saving detailed tests of
models against multi-wavelength luminosity functions and estimates of black hole mass distributions
for future work. However, we are able to draw some interesting general conclusions from our efforts
to reproduce the observed evolution of the optical luminosity function.
At z ≤ 2, we have focused on the Boyle et al. (2000) evolution results, and particularly on their
finding that the QLF has a break that shifts to lower luminosities at lower redshifts. In any model
where p( m) is independent of mass, reproducing this result requires a shift towards lower typical
accretion rates over time (Fig. 5). Decreasing the duty cycle uniformly by reducing the amplitude
of p( m) can lower the normalization of Φ(L), but on its own it cannot shift the break to lower
luminosities, since black hole masses, and thus the location of the break in n(M ), can only increase
with time. Lowering the break luminosity requires decreasing the probability of high accretion
rates relative to low accretion rates. In the context of our double power-law models, this change
is achieved by reducing m∗, from slightly below unity at z ∼ 2 to far below unity at z ∼ 0 − 0.5.
Physically, such a change could arise because of decreasing gas supplies and increasing dynamical
times in galaxies at lower redshifts (Kauffmann & Haehnelt 2000).
A consequence of this reduction in m∗ is a change in the nature of the QLF between high
and low redshift. At high redshift, the sequence of quasar luminosities is primarily a sequence of
black hole masses. Because m∗ is close to unity and p( m) is shallow below the break, most high
luminosity quasars are produced by the more numerous low mass black holes accreting at m ≈ 1
rather than extremely rare high mass black holes with low accretion rates. The typical mass at a
given luminosity is M ≈ L/l, and a factor of ten increase in luminosity roughly corresponds to a
factor of ten increase in black hole mass. However, once m∗ falls well below unity, there is a large
-- 66 --
range m∗ < m < 1 over which p( m) is steep (∝ m−3 for most of our models), thus increasing the
probability that a given luminosity is generated by a high mass black hole accreting at a low m
close to m∗. The typical active black hole mass still increases with luminosity, but at low redshift
a factor of ten increase in L corresponds to only a factor ∼ 3 increase in median black hole mass,
and the range in black hole mass at fixed luminosity is much larger than at high redshift. At low
redshift, the sequence of quasar luminosities remains partly a sequence of black hole mass, but it
is also in large part a sequence of
m (Figs. 14 and 15).
This change in character arises only because we reduce the radiative efficiency in the super-
Eddington regime, forcing systems with m > 1 to radiate at the Eddington luminosity. Without
this transition in the accretion physics, there would be no preferred scale to change the relative
importance of mass and m between high and low redshift, since we assume the same double power
law form of p( m) at all redshifts and it is only the value of
m∗ relative to unity that changes. If
the efficiency does not decrease in the super-Eddington regime, then the shape of the luminosity
function depends in detail on the form and cutoff of p( m) at m > 1. However, the limited data on
masses of high redshift black holes generally shows systems with L/LEdd ≈ 1 but not substantially
larger (McLure & Dunlop 2003; Vestergaard 2004). This result suggests that there is indeed a
break in the radiative efficiency at m ≈ 1, or else that black hole accretion self-regulates to enforce
m . 1, since otherwise it would require p( m) to cut off coincidentally at m > 1 and be relatively
high at m just below one.
If p( m) is mass-dependent, then there is a fundamentally different alternative for explaining
instead of a decline in characteristic accretion
the shift of the QLF break to lower luminosities:
rates, the mass-dependence can itself evolve so that activity is preferentially suppressed in high
mass black holes at low redshift (Fig. 6). Physically, such behavior could arise because high mass
black holes reside in early type galaxies with large bulges, which tend to exhaust their gas supplies
earlier. In this scenario, quasars at a given position on the luminosity function (relative to Lbrk)
are always associated with the same distribution of accretion rates, but the associated black hole
mass declines with redshift as the high mass black holes turn off. In comparison with the mass-
independent p( m) models, the median black hole mass at fixed luminosity is lower, and the range
of black hole masses is smaller, with near-Eddington accretors making a large contribution to the
high end of the luminosity function at all redshifts (see Figure 16). The low black hole mass implies
a high space density of hosts, so this model predicts that a larger fraction of low redshift quasars
reside in late type or low luminosity galaxies.
Based on anecdotal evidence, it is hard to say whether a decline in characteristic accretion rates
or a decline in the activity of high mass black holes is more important in producing the observed
decline of Lbrk at z < 2. The first scenario's prediction of a wider range of L/LEdd at lower redshifts
seems in qualitative agreement with studies of black hole masses (Woo & Urry 2002; Vestergaard
2004). The second scenario seems in qualitative agreement with the relative quiescence of the black
holes in most massive ellipticals (such as M87), though Kauffmann et al. (2003) find that the most
luminous AGN in the local universe do reside in massive, early type systems. The two scenarios
-- 67 --
make quantitatively different predictions, and careful comparison to studies of active black hole
masses and host galaxy properties should be able to show whether one mechanism dominates or
both are comparably important. The results of Vestergaard (2004) suggest that active black hole
masses at low redshift are intermediate between the values predicted by the two models shown in
Figure 16.
At high redshifts, constraints on the form of the QLF are weaker; in particular, the SDSS
measurements of Fan et al. (2001) probe only the high luminosity end of Φ(L). For matching the
observed evolution over the range z ∼ 5 to z ∼ 2, we find one acceptable solution in which p( m) is
roughly constant in Eddington units and the growth of the black holes themselves drives the growth
in amplitude of the QLF (Fig. 7). With our adopted parameters, the black holes grow by a factor
∼ 10 between z = 5 and z = 2, and the space density at z = 2 is n∗M∗ = 2.012 × 10−5 Mpc−3 at
M∗ = 109M⊙. Significantly lower normalizations of n(M ) at z = 2 are not allowed because they
would require a negative black hole mass density at z = 5. Higher normalizations are allowed, in
which case the quasar duty cycle is shorter, the rate of black hole growth is smaller, and the growth
of the QLF is driven by a steady increase in p( m). While a solution with a high black hole space
density can give an acceptable match to the QLF, it seems physically unattractive because it requires
that most of the black hole mass density was already in place at z = 5, before the main epoch of
quasar activity. In this latter scenario, the luminous phases of quasars represent the addition of a
small amount of mass to already formed black holes. If we assume instead that the observed optical
QLF does trace the growth of black holes, and that the former model is therefore more realistic,
then our analysis predicts a black hole space density M n(M ) ∼ 2 − 3 × 10−5 Mpc−3 at M = 109M⊙
at z = 2, and continuation to z = 0 implies a similar space density at M ∼ 2 × 109M⊙. However,
we have not investigated the sensitivity of this prediction to our specific choices of parameters,
such as the double power-law form and adopted slopes of p( m) and n(M ). Black hole mergers and
obscured accretion could also alter the prediction significantly, especially at low redshift.
6.3. Distinguishing Scenarios
Many of the qualitative results mentioned above could have been anticipated without detailed
calculations. Our framework, however, allows one to compute quantitative predictions of concrete
models that illustrate distinct ideas about the nature of black hole and quasar evolution. We did this
in §5 for the low redshift (z ≤ 2) regime, adopting as our baseline a model with quasar emission and
black hole growth dominated by unobscured thin-disk accretion and a normalization of n(M, z = 2)
implying a short quasar lifetime, and, consequently, little growth of black hole masses from z = 2 to
z = 0. We compared this model to four variants: one with a lower n(M, z = 2) and correspondingly
longer quasar lifetime, one with a 4:1 ratio of obscured to unobscured systems, one with a large
amount of merger driven growth of black hole masses, and one with a boosted probability of low- m
accretion leading to substantial ADAF growth of black holes. For each scenario, we are able to find
parameters that acceptably reproduce the Boyle et al. (2000) optical QLF at z = 2, 1, and 0.5.
-- 68 --
Relative to the short-tq model, the long-tq model starts at z = 2 with a factor ∼ 6 lower n(M )
at every mass. Because of the larger amount of accretion per black hole, however, the long-tq n(M )
overtakes the short-tq n(M ) at high masses by z = 0, while remaining below it at low masses. The
two models have similar QLFs at every wavelength, since they match in the optical by construction
and are dominated by systems with the Elvis et al. (1994) SED. However, the growth of M∗ in the
long-tq model requires a more rapid decline of
m∗ to compensate, so at lower redshifts it predicts
lower median m and higher median black hole mass at a given luminosity. The two models can thus
be distinguished observationally by the z = 0 black hole mass function, by the mass distributions
of active black holes at z ∼ 0.5 − 1, and by the space densities of host systems, which are lower in
the long-tq model at every luminosity and redshift.
In the obscured model, the large amount of obscured accretion produces more black hole growth
than in the short-tq model, leading to a higher n(M ) and ρbh at low redshift. By z = 0, n(M ) is
higher by a factor ∼ 6 at high masses. However, the median black hole mass and host space density
at a given optical luminosity are only slightly higher. The clearest distinguishing feature of this
model is the relative amplitude of luminosity functions in different wavelength bands, a consequence
of the different SED shapes of obscured and unobscured accretors. At z = 2, the 2-10 keV luminosity
function is elevated by nearly a factor of five, the ratio of all systems to unobscured systems. This
boost decreases towards low redshift because of the increasing importance of obscuration in the
(observed-frame) 2-10 keV band. The 0.5-2 keV band is heavily obscured at z < 2, so the soft X-
ray luminosity function of the obscured model is similar to that of the short-tq and long-tq models,
except at z = 2 where it has a higher amplitude at low luminosity. The strongest departure of all is
in the FIR, where the re-radiated emission of obscured accretors boosts Φ(L) by factors of ten (low
luminosity) to one hundred (high luminosity), relative to the short-tq and long-tq models. This
crucial prediction should soon be testable by SIRTF and by other sub-mm and mm-wavelength
observations.
In the merger model, low redshift mergers strongly distort the initial black hole mass function,
depleting it at low masses and boosting it at high masses, with a factor of 16 increase in the high
mass end at z = 0 relative to the short-tq model. The quasar population of this model is still
dominated by systems with a thin-disk SED, and since it matches the observed optical luminosity
function by construction, its predictions at other wavelengths are close to those of the short-tq
and long-tq models. However, the high space density of high mass black holes leads to a high
median mass of active black holes at fixed luminosity, similar to that of long-tq at low luminosities
and higher still at high luminosities. Merger driven distortions of the mass function also lead to
distinctive redshift and luminosity dependence of the host space density.
Finding parameters that yield significant ADAF growth and an acceptable match to the optical
luminosity function proves quite difficult, requiring an artificial boost to the probability of accretion
rates below mcrit. With our adopted parameters, the ADAF model predicts a large amount of
black hole growth between z = 2 and z = 0, and thus a high n(M ) and ρbh at z = 0. With the
combination of high M∗ and high probability of low m, the ADAF model predicts the largest median
-- 69 --
black hole masses and lowest median accretion rates at fixed luminosity, with the median accretion
rate approaching mcrit even for high luminosity AGN at z ≤ 0.5. The high X-ray fraction of the
ADAF SED boosts the soft and hard X-ray luminosity functions relative to the optical, especially at
low redshift, and the model predicts that a majority of X-ray selected systems at z ∼ 0.5 should be
ADAF accretors, even at high luminosities. This prediction appears observationally untenable, and
the model requires rather implausible parameter choices in the first place, so our results suggest that
ADAFs are unlikely to make an important contribution to black hole growth in the real universe,
even at z < 2 (Haehnelt et al. [1998] reach a similar conclusion). Low radiative efficiency at low m
may nonetheless help explain the remarkable quiescence of most black holes in the local universe
(Narayan et al. 1998).
6.4. Prospects
Our results illustrate how a variety of observational constraints can be brought to bear on
the key questions of quasar and black hole evolution. In particular, we have extended the ideas of
Soltan (1982) and Small & Blandford (1992) to show that incorporating the link between luminous
accretion and black hole growth allows one to construct concrete physical models that are simulta-
neously constrained by multi-wavelength luminosity function measurements and estimates of black
hole masses and accretion rates. For our models in §5, we chose a plausible but not unique set of
initial conditions at z = 2 and evolved them forward in time under varying assumptions, always
matching the observed optical QLF. The models then make distinguishable predictions for other
observables.
With our current parameter choices, all of our models face some difficulty when confronted
with recent estimates of X-ray luminosity functions and the local black hole mass function, as
illustrated in Figures 10, 12, and 13. Models that fit the Ueda et al. (2003) hard X-ray QLF
generally do not fit the Miyaji, Hasinger, & Schmidt (2001) soft X-ray QLF at the same redshift,
and vice versa. A model incorporating luminosity and redshift dependence of the obscured quasar
fraction might fare better, though some of the problem may still lie with the observational estimates
themselves. Combining Tremaine et al.'s (2002) estimate of the M − σ relation with Sheth et al.'s
(2003) estimate of the distribution of galaxy velocity dispersions yields a mass function that lies
well below our model predictions for M > 109M⊙, unless the intrinsic scatter of the M − σ relation
is ∼ 0.5 dex, compared to Tremaine et al.'s estimate of ≤ 0.3 dex. Repairing this discrepancy would
require either reducing the break mass at z = 2 substantially below M∗ = 109M⊙ or dropping our
assumed double power-law form of n(M ) and adopting a mass-dependent p( m) to reproduce the
Boyle et al. (2000) QLF. For the present, we do not want to draw strong conclusions from these
discrepancies, since we have not thoroughly assessed the observational uncertainties, and we have
not investigated the extent to which a failing model can be "fixed up" by adjusting its parameters
(e.g., the initial shape of the black hole mass function), while retaining its essential features (e.g.,
a large fraction of obscured systems).
-- 70 --
The z = 0 black hole mass function can be reasonably well estimated from current data, at
least at masses M ≤ 109M⊙ where the form and scatter of the M − σ relation are well constrained,
and it is a fundamental boundary constraint on any evolution model. For a comprehensive attempt
to match observations, therefore, it probably makes sense to impose this constraint a priori on all
models, and integrate the evolutionary equations backward in time. In our framework, this approach
is just as easy as integrating forward, even if it is less intuitive. In effect, one takes the known black
hole mass function today, infers the accretion rate distribution by matching the luminosity function,
steps backward by removing the implied amount of mass from each bin of the mass function, and
repeats. We will apply this approach to available observations in future work. Given the inevitable
uncertainties in the observational data, radiative efficiencies, and bolometric corrections, there
are likely to be some degeneracies in the solutions, but we can hope that models that differ in
fundamental rather than incidental features will remain observationally distinguishable. Based
on the results found here, we suspect that the primary source of uncertainty will be the mass
dependence of p( mM ), which requires accurate measurements of both the QLF and the masses
of active black holes to pin down empirically. Mergers look like the other most difficult problem,
though in this case there are good theoretical ideas about what the merger rates of dark halos
and galaxies should be (e.g., Taylor & Babul 2001), and these can be incorporated into model
calculations.
The traditional picture of quasars as a population of supermassive black holes growing by
accretion seems more secure than ever. Many open questions remain about the roles of black
hole mass, accretion rate, radiative efficiency, and SED shape in determining quasar luminosities,
about the properties of accretion flows at low and high accretion rates, about the importance of
black hole mergers and obscured accretion as drivers of black hole growth, and about the relations
among populations observed at different wavelengths. Our work highlights a number of areas
where observational advances will be crucial to answering these questions. These include improved
determination of the local black hole mass function, better understanding of the dependence of
radiative efficiency and SED shape on accretion rate, measurements of the luminosity function
at different wavelengths over the widest achievable range in luminosity and redshift, estimates of
masses and accretion rates of active black holes as a function of redshift and luminosity, and indirect
estimates of black hole space densities from host galaxy and quasar clustering studies. Fortunately,
the observational situation is advancing rapidly, and many of these areas have seen substantial
progress in the last few months alone, as discussed in §5.
It is worth emphasizing the value of
luminosity function determinations and black hole mass estimates that traverse the break in the
QLF and extend as far below as possible. Accurate characterization of this regime is crucial for
separating the roles of n(M ) and p( m) in shaping the luminosity function, which in turn is necessary
for understanding the contribution of sub-Eddington accretion rates to black hole mass evolution.
These lower luminosities are also where optical and X-ray evolution appear to be radically different,
and better measurements of the joint X-ray, optical, and IR luminosity functions are needed to pin
down the origin of these differences. The emerging data on black hole and quasar evolution are
complex, complementary, and rich. We hope that the physical modeling approach described in this
paper will prove useful in exploiting their power.
-- 71 --
We thank Jordi Miralda-Escud´e for crucial suggestions during the early stages of this work.
We also thank Jordi and Martin Haehnelt, Paul Martini, and Smita Mathur for providing detailed
comments on the manuscript that led to significant improvements. We have also benefited from
discussions with many other colleagues, including Ramesh Narayan, Patrick Osmer, Bradley Peter-
son, Richard Pogge, Hans-Walter Rix, and Marianne Vestergaard. The discussion of quasar hosts
in §5.5 was stimulated by conversations with Hans-Walter Rix. This work was supported by NSF
grants AST-0098515 and AST-0098584.
-- 72 --
REFERENCES
Abramowicz, M. A., Czerny, B., Lasota, J. P., & Szuszkiewicz, E. 1988, ApJ, 332, 646
Aller, M. C. & Richstone, D. 2002, AJ, 124, 3035
Anderson, S.F., et al. 2003, AJ, in press, astro-ph/0305093
Barger, A. J., Cowie, L. L., Bautz, M. W., Brandt, W. N., Garmire, G. P., Hornschemeier, A. E.,
Ivison, R. J., & Owen, F. N. 2001, AJ, 122, 2177
Barger, A. J., Cowie, L. L., Brandt, W. N., Capak, P., Garmire, G. P., Hornschemeier, A. E.,
Steffen, A. T., & Wehner, E. H. 2002, AJ, 124, 1839
Begelman, M. C. 1978, MNRAS, 243, 610
Begelman, M. 2002, ApJ, 568, L97
Blandford, R. D., & Begelman, M. C. 1999, MNRAS, 303, 1
Brandt, W.N., et al. 2001, AJ, 122, 281
Boyle, B. J., Shanks, T., Croom, S. M., Smith, R. J., Miller, L., Loaring, N., & Heymans, C. 2000,
MNRAS, 317, 1014
Cavaliere, A., & Vittorini, V. 2000, ApJ, 543, 599
Cavaliere, A., & Vittorini, V. 2002, ApJ, 570, 114
Chokshi, A., & Turner, E. L. 1992, MNRAS, 259, 421
Comastri, A., Fiore, F., Vignali, C., Matt, G., Perola, G.C., LaFranca, F. 2001, MNRAS, 327, 781
Comastri, A., Setti, G., Zamorani, G., & Hasinger, G. 1995, A&A, 296, 1
Corbett, E. A., et al 2003, MNRAS, 343,705
Cowie, L. L., Barger, A. J., Bautz, M. W., Brandt, W. N., & Garmire, G. P. 2003, ApJ, 584, 57
Croom, S. et al. 2003, to appear in AGN Physics with the Sloan Digital Sky Survey, ed. G.T.
Richards & P.B. Hall (San Francisco:ASP), astro-ph/0310533
Czerny, B., Niko lajuk, M., Piasecki, M., & Kuraszkiewicz, J. 2001, MNRAS, 325, 865
Czerny, B., Nikolajuk, M., Rozanska, A., Dumont, A. M., Loska, Z., & Zycki, P.T. 2003, A&A, in
press, astro-ph/0309242
Dunlop, J. S., McLure, R. J., Kukula, M. J., Baum, S. A., O'Dea, C. P., Hughes, D. H. 2003,
MNRAS, 340, 1095
-- 73 --
Elvis, M., Risaliti, G., & Zamorani, G. 2002, ApJ, 565, 75
Elvis, M., Wilkes, B. J., McDowell, J. C., Green, R. F., Bechtold, J., Willner, S. P., Oey, M. S.,
Polomski, E., & Cutri, R. 1994, ApJS, 95, 1
Fan, X., et al. 2001, AJ, 121, 54
Efstathiou, G., & Rees, M. J. 1988, MNRAS, 230, 5
Esin, A. A., McClintock, J. E., & Narayan, R. 1997, ApJ, 489, 865
Fabian, A. C. 2003, to appear in Carnegie Observatories Astrophysics Series, Vol. 1: Coevolution
of Black Holes and Galaxies, ed. L. C. Ho (Cambridge: Cambridge Univ. Press), astro-
ph/0304122.
Fabian, A. C., & Iwasawa, K. 1999, MNRAS, 303, 34
Fiore, F., La Franca, F., Giommi, P., Elvis, M., Matt, G., Comastri, A., Molendi, S., & Gioia, I.
1999, MNRAS, 306, 55
Fiore, F. et al. 2003, A&A, 409, 79
Gebhardt, K., et al. 2000, ApJ, 539, 13
Gebhardt, K., et al. 2000, ApJ, 543, 5
Giacconi, R., et al. 2002, ApJS, 139, 369
Gilli, R., Salvati, M., & Hasinger, G. 2001, A&A, 366, 407
Haenhelt, M. G., Natarajan, P., & Rees, M. J. 1998, MNRAS, 300, 817
Haenhelt, M. G., & Rees, M. J. 1993, MNRAS, 263, 168
Haiman, Z., & Loeb, A. 2001, ApJ, 552, 459
Haiman, Z., & Hui, L. 2001, ApJ, 547, 27
Hasinger, G. 2003, to appear in High Energy Processes and Phenomena in Astrophysics, IAU
Symposium 214, eds. X. Li, Z. Wang, & V. Trimble, astro-ph/0301040
Ho, L. 2001, in IAU Colloq. 184, AGN Surveys, eds. R. F. Green, E. Ye. Khachikian, & D. B.
Sanders (San Francisco: ASP), astro-ph/0110438
Ho, L. 1999, ApJ, 516, 672
Kaiser, N. 1984, ApJ, 284, 9
Katz, J. 1977, ApJ, 215, 265
-- 74 --
Kauffmann, G., & Haehnelt, M. 2000 MNRAS, 311, 576
Kauffmann, G. & Haehnelt, M. G. 2002, MNRAS, 332, 529
Kauffmann, G., et al. 2003, MNRAS, submitted, astro-ph/0304239
Kuraszkiewicz, J. K., Wilkes, B. J., Czerny, B., Mathur, S., Brandt, W. N., & Vestergaard, M.
2000, New Astronomy Review, 44, 573
Laor, A. 1998, ApJ, 505, L83
Lee, M. H. 2000, Icarus, 143, 74
Lynden-Bell, D. 1969 Nature, 223, 690
Madau, P., Rees, M. J., Volonteri, M., Haardt, F., & Oh, S. P. 2003, ApJ, submitted, astro-
ph/0310223
Martini, P., & Weinberg, D. H. 2001, ApJ, 547, 12
Martini, P. 2003, to appear in Carnegie Observatories Astrophysics Series, Vol. 1: Coevolution
of Black Holes and Galaxies, ed. L. C. Ho (Cambridge: Cambridge Univ. Press), astro-
ph/0304009
McLure, R. J., & Dunlop, J.S. 2002, MNRAS, 331, 795
McLure, R. J., & Dunlop, J.S. 2003, MNRAS, submitted, astro-ph/0310267
Menou, K., Haiman, Z., & Narayanan, V. K. 2001, ApJ, 558, 535
Merritt, D., & Ferrarese, L. 2000, MNRAS, 320, 30
Merritt, D. & Ferrarese, L. 2001, MNRAS, 320, L30
Miyaji, T., Hasinger, G., & Schmidt, M. 2001, A&A, 353, 25
Mo, H.J., & White S.D.M. 1996, MNRAS, 282, 1096
Murali, C., Katz, N., Hernquist, L., Weinberg, D. H., & Dave, R. 2002, ApJ, 571, 1
Murray, S. D., & Lin, Douglas N. C. 1996, ApJ, 467, 265
Narayan, R., Mahadevan, R., & Quataert, E. 1998, in Theory of Black Hole Accretion Disks, eds.
Marek A. Abramowicz, Gunnlaugur Bjornsson, & James E. Pringle. (Cambridge: Cambridge
University Press), p. 148, astro-ph/9803141
Onken, C.A., & Peterson, B.M. 2002, ApJ, 585, 121
Priddey, R. S., Isaak, K. G., McMahon, R. G., & Omont, A. 2003, MNRAS, 339, 1183
-- 75 --
Pei, Y. C. 1995, ApJ, 438, 623
Quataert, E., di Matteo, T., Narayan, R., Ho, L. 1999, ApJ, 525, 89
Quataert, E., & Gruzinov, A. 2000, ApJ, 539, 809
Redmount, I. H., & Rees, M. J. 1989, Com Ap, 14, 165
Rees, M. J. 1978, The Observatory, 98, 210
Richstone, D., et al. 1998 Nature, 395, 14
Salpeter, E. E. 1964, ApJ, 140, 796
Salucci, P., Szuszkiewicz, E., Monaco, P., & Danese, L. 1999, MNRAS, 307, 637
Sazonov, S. Yu., Ostriker, J.P., & Sunyaev, R.A. 2003, MNRAS, submitted, astro-ph/0305233
Schmidt, M. 1968, ApJ, 151, 393
Schmidt, M., Schneider, D.P., & Gunn, J.E. 1995, AJ, 107, 1245
Schneider, D. P., et al. 2002, AJ, 123,567
Setti, G. & Woltjer, L. 1989, A&A, 224, L21
Silk, J., Rees, M. J. 1998, A&A, 331, 1
Sheth, R. K. 1998, MNRAS, 295, 869
Sheth, R., et al. 2003, ApJ, 594, 225
Silk, J. & Takahashi, T. 1979, ApJ, 229, 242
Silk, J. & White, S. D. 1978, ApJ, 223, L59
Small, T. A., & Blandford, R. D. 1992, MNRAS, 259, 725
Soltan, A. 1982, MNRAS, 200, 115
Steffen, A. T., Barger, A. J., Cowie, L. L., Mushotzky, R. F., & Yang, Y. 2003, ApJ, 596, 23
Steidel, C. C., Hunt, M. P., Shapley, A. E., Adelberger, K. L., Pettini, M., Dickinson, M., Giavalisco,
M. 2002, ApJ, 576, 653
Taylor, J.E., & Babul, A. 2001, ApJ, 559, 716
Tremaine, S., et al. 2002, ApJ, 574, 740
Ueda, Y., Akiyama, M., Ohta, K., & Miyaji, T. 2003, ApJ, in press, astro-ph/0308140
-- 76 --
Valtonen, M. J., Mikkola, S., Heinamaki, P., & Valtonen, H. 1994, ApJS, 95, 69
Vestergaard, M. 2004, ApJ, 600, in press, astro-ph/0309521
Wandel, A., Peterson, B. M., & Malkan, M.A. 1999, ApJ, 526, 579
Warren, S.J., Hewett, P.C., & Osmer, P.S. 1994, ApJ, 421, 412
White, R.L., et al. 2000, ApJS, 126, 133
Wilkes, B. J., Schmidt, G. D., Cutri, R. M., Ghosh, H., Hines, D. C., Nelson, B., & Smith, P. S.
2002, ApJ, 564, 65
Wisotzki, L. 2000, A&A, 353, 853
Wolf, C., Wisotzki, L., Borch, A., Dye, S., Kleinheinrich, & Meisenheimer, K. 2003, A&A, 408, 499
Woo, J-H., & Urry, C. M. 2002, ApJ, 579, 530
Wyithe, J. S. B., & Loeb, A. 2003, ApJ, 595, 614
Yu, Q. & Tremaine, S. 2002, MNRAS, 335, 965
This preprint was prepared with the AAS LATEX macros v5.0.
-- 77 --
A. Luminosity Function for a Double Power-Law p( m)
For the double power-law p( m) and double power-law n(M ), the convolution integral (8) for
the luminosity function must be broken into three different regimes to account for the different
efficiencies of the accretion modes. The total QLF is the sum of the QLFs produced by each
accretion mode, Φ(L)Total = Φ(L)SE + Φ(L)TD + Φ(L)ADAF. The calculation is analogous to that
in §3.1, though more tedious, and we omit the details. The solution depends on whether m∗ lies
in the thin-disk, super-Eddington, or ADAF regimes. For the first case, which is the one usually
relevant to our models, the results for the three accretion modes are
+ 1− mb−α
(b−α) mb
∗
l(b+1) mb
l(b+1) mb
∗
∗
l
l
n∗p∗
n∗p∗
(a−α) ma
∗ −0.01a−α
Φ(L)SE =
h ma−α
lM∗(cid:17)a
h (β−α) m−a
(a−α)(a−β) (cid:16) L
lM∗(cid:17)α
∗ (cid:16) L
l (cid:20) (β−α) m−b
lM∗(cid:17)b
(b−α)(b−β) (cid:16) L
lM∗(cid:17)α
∗ (cid:16) L
+ 1
+ 1
(b−α) mb
(b−α) mb
∗
n∗p∗
n∗p∗
n∗p∗
∗
+ (b−a) m−α
lM∗(cid:17)α
∗ (cid:2)10b+1 − 1(cid:3)(cid:16) L
lM∗(cid:17)β
∗ (cid:2)10b+1 − 1(cid:3)(cid:16) L
lM∗(cid:17)α
∗i(cid:16) L
lM∗(cid:17)α
(a−α)(b−α) (cid:16) L
lM∗(cid:17)β(cid:21)
∗ (cid:16) L
lM∗(cid:17)β
(a−β)(b−β)(cid:16) L
lM∗(cid:17)β(cid:21)
∗ (cid:16) L
lM∗(cid:17)β
∗i(cid:16) L
+ (b−a) m−β
∗
− 0.01a−β
(a−β) ma
− 0.01a−β
(a−β) ma
n∗p∗
l
∗ −0.01a−β
(a−α) ma
∗
+ 1− mb−β
(b−α) mb
∗
h ma−β
(A1)
(A2)
L < lM∗
L > lM∗ ,
L < 0.01lM∗
0.01lM∗ < L < m∗lM∗
m∗lM∗ < L < lM∗
L > lM∗ ,
Φ(L)TD =
and
Φ(L)ADAF =
n∗p∗0.01α+1
l(a−2α−1) ma
n∗p∗
l ma
∗ (cid:20)
2
a−1
2(β−α)
lM∗(cid:17)
(a−2β−1)(a−2α−1) (cid:16) L
lM∗(cid:17)β
(cid:16) L
lM∗(cid:17)α
∗ (cid:2)0.01a−2α−1 − 10−4(a−2α−1)(cid:3)(cid:16) L
lM∗(cid:17)α(cid:21)
a−2α−1 (cid:16) L
lM∗(cid:17)β
∗ (cid:2)0.01a−2β−1 − 10−4(a−2β−1)(cid:3)(cid:16) L
+ 0.01a−2α−1
− 10−4(a−2β−1)
a−2β−1
n∗p∗0.01β+1
l(a−2β−1) ma
L < 10−4lM∗
10−4lM∗ < L < 0.01lM∗
L > 0.01lM∗ ,
(A3)
mmax = 10, and mmin = 10−4 are explicitly included in the solutions.
where the values mcrit = 0.01,
The three regimes differ because of the different dependence of ǫ0.1 on m and because m∗ lies in
the thin-disk regime, breaking that integral into more parts. Note that the slopes at the high and
low luminosity end of each mode are equal to the mass function slopes (α and β), for the reasons
discussed in §3.1.
-- 78 --
B. Critical Slopes
We gain some insight into results for general p( m) and n(M ) by considering cases in which
p( m) = ma is a pure power-law in some range mmin − mmax and n(M ) = M α is a pure power-law
in some range Mmin − Mmax. Analysis of such cases reveals a number of critical slopes where the
character of the solutions changes. For the critical p( m) slope a = −2, each logarithmic range
of
m contributes equally to black hole growth and, if ǫ0.1 is constant, to emissivity of the quasar
population. When a ≪ −2, growth and emissivity are dominated by low accretion rates (near
mmin), and when a ≫ −2 high accretion rates (near mmax) dominate. Our double power-law
models have a low- m slope a > −2 and a high- m slope b < −2, so that the largest contributions
to growth and emissivity are from accretion rates near m∗, which we usually take (at least at high
redshift) to lie in the thin-disk range 0.01 < m∗ < 1. These parameter choices make our results
relatively insensitive to our assumptions about efficiencies and the form of p( m) in the ADAF
and super-Eddington regimes. However, low- m or high- m accretion may be more important in
the real universe, at least at some redshifts, in which case the form of p( m) and behavior of the
accretion physics in these regimes would have a larger impact on observable properties of the quasar
population.
The distribution of active black hole masses at a fixed luminosity depends on both the p( m)
and n(M ) slopes. When α < a+ 1 and ǫ0.1 = constant, the black hole mass function is steep enough
that low mass black holes with high accretion rates predominate (eq. 21) -- i.e., the most common
black holes at a given luminosity L are either those with M = Mmin or those with the maximum
accretion rate and M = L/(ǫ0.1l mmax). Conversely, when α > a + 1, high mass black holes with low
accretion rates predominate. For more general forms of n(M ) and p( m), the roles of "minimum"
and "maximum" values are in practice played by values where the slope of the distribution changes
or where there is a break in efficiency. For example, if α < a + 1 and the power-law behavior of
p( m) extends to m ≈ 1, then decreasing efficiency in the super-Eddington regime comes into play,
and the luminosity function is dominated by black holes radiating near the Eddington luminosity.
This is the typical behavior for our double power-law models at high redshift, where m∗ is close to
unity and the low- m slope of p( m) and high mass slope of n(M ) easily satisfy β < a + 1. At low
redshift we have low values of
m∗, so the high- m slope of p( m) becomes important. Even here we
usually have β < b + 1, so that systems with m ≈ m∗ dominate the high luminosity end of Φ(L),
but because the slopes that we adopt (b = −3, β = −3.4) are not so far from the critical relation,
systems at a given luminosity span a wide range of black hole masses and accretion rates.
Two more critical slopes arise when we ask whether the space density of black holes of mass
M increases or decreases with time. If accretion drives the evolution of n(M ), then the critical
slope is −1: for a mass function steeper than n(M ) ∝ M −1, accretion increases n(M ), while for a
shallower slope the number of black holes "lost" to higher masses exceeds the number gained from
lower masses, driving n(M ) down with time. For merger driven growth, the corresponding critical
slope is −2, at least if objects merge with others of equal mass as in the simple models considered
here. The difference in slopes arises because accretion adds mass to the black hole population
-- 79 --
while mergers do not, and the critical slope for mergers is steeper still if black hole ejection or
gravitational radiation reduce the mass of the surviving merger product below the combined mass
of its progenitors. For black hole mass functions whose asymptotic slopes match the asymptotic
slopes of the Boyle et al. (2000) luminosity function, mergers drive n(M ) up in the high mass regime
and down in the low mass regime. Accretion increases n(M ) in both regimes, but the increase is
slow for low masses and rapid for high masses.
|
astro-ph/0111557 | 1 | 0111 | 2001-11-29T14:03:26 | Time-dependent correlations of inflationary perturbations | [
"astro-ph",
"gr-qc",
"hep-ph",
"quant-ph"
] | We show that if the primordial classical perturbations were generated by the gravitational particle creation during inflation, and followed by an evolution of quantum-to-classical transition,the time dependent correlation of these perturbations is long-tailed with a correlation time larger than the Hubble-time. Consequently, the inflationary perturbations are locally scale-scale correlated. Hence, the interaction of the fields during inflation can be explored via the detection of the local scale-scale correlation of the CMB fluctuations. | astro-ph | astro-ph | Europhysics Letters
PREPRINT
1
0
0
2
v
o
N
9
2
1
v
7
5
5
1
1
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Time dependent correlations of inflationary perturbations
W.-L. Lee 1 and L.-Z. Fang 2
1 Institute of Physics, Academia Sinica, Taipei, Taiwan 11529, Republic of China
2 Department of Physics, University of Arizona, Tucson, AZ 85721, U.S.A.
PACS. 98.70.Vc -- Background radiations.
PACS. 98.80.Cq -- Inflationary universe.
Abstract. --
We show that if the primordial classical perturbations were generated by the gravitational
particle creation during inflation, and followed by an evolution of quantum-to-classical tran-
sition, the time dependent correlation of these perturbations is long-tailed with a correlation
time larger than the Hubble-time. Consequently, the inflationary perturbations are locally
scale-scale correlated. Hence, the interaction of the fields during inflation can be explored via
the detection of the local scale-scale correlation of the CMB fluctuations.
It is fundamentally important to understand the nature of the primordial cosmological
perturbations which are responsible for the formation of structures that we see today. The
inflation paradigm for the early universe assumes that the cosmic perturbations are initiated
by the process of particle creation from vacuum in the background gravitational field of the
expanding universe [1]. The subsequent decoherence of the quantum fluctuations leads to a
quantum-to-classical transition [2], and gives rise to the initial perturbations in the radia-
tion dominated era on scales beyond the Hubble radius. Therefore, the earliest evolution of
the inflationary cosmological fluctuations is governed by the gravitational particle creation
accompanying the quantum-to-classical transition. Recently, the mechanism of the gravita-
tional particle creation is re-emphasized with the development of the quintessential inflation
model [3]. However, it still seems to lack some testable predictions of these mechanisms.
In this Letter, we will show that the time-dependent correlation of the primordial perturba-
tions is one of such testable predictions. In the "standard" inflation, i.e. the slow-roll inflation
caused by a single scalar inflaton φ, the correlation time of the primordial perturbations is
longer than the Hubble time H −1. This correlation will entail a non-trivial observable feature
-- the local scale-scale correlation of the initial perturbations.
The time-dependence of cosmological perturbations is generally not observable in non-
inflationary models, because the observed mass field does not contain two or multi-time in-
formation of the perturbations. In contrast, the inflationary scenario provides an excellent
prospect to observing the time-dependence of the initial perturbations in a given region on
the scale of horizon. Owing to the so-called "first out -- last in" feature, a fluctuation crossing
over the Hubble radius H −1 at the moment t× will yield a perturbation at the end of the
inflation, tf , with the physical wavenumber given by [4]
kp ∝ eH(t×−tf ).
(1)
c(cid:13) EDP Sciences
2
EUROPHYSICS LETTERS
1 and t×
Consequently, the longer the physical length of the perturbations, the earlier the time of be-
coming the super-horizon scaled, and vice versa. Therefore, if perturbations with the same
physical scale at two different instants t×
2 are correlated during the super-horizon evo-
lution, the perturbations of the corresponding physical scales kp(t×
2 ) at the same
time tf will be correlated. Equation (1) prescribes exactly the mapping of the correlation be-
tween two perturbations on the same physical scale with different horizon-crossing moments
to that of two perturbations at the same time with different physical scales. The latter is
actually observable. If the length scales of perturbations are larger than the Hubble radius
H −1 at the decoupling era, the perturbations were less contaminated by the post inflationary
reheating and other small scale processes. Thus, one can expect that the two-scale correla-
tion of the cosmic background temperature fluctuations retain the information of the time
dependent behavior of the initial perturbations.
1 ) and kp(t×
The time-dependent correlation is not a new issue. It is generally believed that the per-
turbations originated from the quantum fluctuations of vacua during inflation have large cor-
relation time since they vary slowly in the super-horizon regime. However, the rough idea of
the lengthy correlation time is inadequate to establish the detectability of the time-dependent
correlation , and one must study the temporal correlation of the perturbations in both quan-
tum and classical regime to reach the equation suitable of calculating the time correlation in
practical.
Let us calculate the time-dependent behavior for the "standard" slow-roll inflation gov-
erned by a real massless scalar field φ described by the action
S =
1
2Z d4x√−g[∂µφ∂µφ − V (φ)],
(2)
where V (φ) is a self-interaction potential. Under the slow-roll condition, the time-scale of the
self-interaction is much longer than 1/H. That is, the self-interaction is negligible if we study
only the evolution on time scales smaller than the duration of inflation. Without interaction,
the evolution of perturbations of the free φ field mode will be coherent. Thus, the φ field
self-interaction is ineffective of violating the temporal coherence during inflation.
For the simple model Eq. (2), the dominant interaction during inflation is the coupling
between the φ field and the gravitation of the expanding universe, governed by the gravita-
tional scalar density √−g in the integral. Therefore, in the "standard" inflation model, the
origin and evolution of the temporal coherence of the primordial perturbations are actually
determined by the dynamics of the gravitationally driven particle creation.
In a de Sitter space, the scalar field φ(r, t) can be described as a superposition of free modes
with coordinate and conjugate momentum variables given by φ(k) = (2π)−3/2R φ(r)e−ik·rd3r
and π(k) = a2φ′(k), respectively, where k is the comoving wavevector and a = eHt represents
the cosmic scale factor [5]. The notation ′ denotes the derivative with respect to the conformal
time τ =R dt/a(t) = −(1/H)e−Ht. Therefore, the time-dependent behavior of the φ field can
be studied via the modes k.
given by [φ(k, τ ), π†(k, τ )] = iδ(3)(k − k′), which yields
The quantum nature of the φ field is specified by the equal-time commutation relation
φ(k, τ ) = σk(τ )ak + σ∗
π(k, τ ) = a2σ′
k(τ )a†
k(τ )ak + a2σ∗′
−k,
k (τ )a†
−k.
(3)
The ak and a†
k are respectively, the annihilation and creation operators, satisfying commuta-
W.-L. Lee and L.-Z. Fang: Time dependent correlations of inflationary perturbations3
tion relations [ak, a†
k′] = δ(3)(k − k′). The time dependence of the mode k is given by
σk(τ ) =
1
√2k (cid:20)Hτ − i
H
k (cid:21) e−ikτ ,
If the system is in the vacuum state at a given time τ0, we have
ak0, τ0i = 0
(4)
(5)
for all k. In the Schrodinger representation, the evolution of the system leads the vacuum
state at τ0 developing into 0, τi = S0, τ0i at a later time τ , where S is the S-matrix. The
state 0, τi is no longer a vacuum state. The average number of the particles gravitationally
created in mode k is Nk(τ ) = h0, τ0S−1a†
One cannot directly calculate the time-dependent correlation of classical density pertur-
bations by the state 0, τi, because at a different τ the states are quantum coherent. We
should first find the condition of quantum decoherence for the system at different times. In
the coordinate representation φ(k), the Schrodinger wave function of the state 0, τi is given
by the product of the wave functions for all k modes [6]:
ka−kS0, τ0i.
Ψ[φ(k), τ ] =
1
√πσk(τ )
exp(cid:26)− φ(k)2
2σk(τ )2 [1 − i2F (k, τ )](cid:27) ,
where
F (k, τ ) =
1
2kτ
H
2k
= −
eHt.
(6)
(7)
Equation (6) is actually the ground state wave function of a harmonic oscillator with coordi-
nate φ(k) which possesses a time-dependent variance in itself.
as
Thus, the quantum coherence between the coordinates at different times can be calculated
hΨ(τ2)φ†(k)φ(k)Ψ(τ1)i =
1
2√πσk(τ1)σk(τ2)
1
(cid:26)
2σk(τ1)2 [1 − i2F (k, τ1)] +
1
2σk(τ2)2 [1 + i2F (k, τ2)](cid:27)−3/2
(8)
.
When t ≫ H −1 or Hτ = e−Ht ≪ 1, the term Hτ in Eq. (4) can be ignored, and σk(τ )
becomes τ -independent. Accordingly, σk(τ1) ≃ σk(τ2), and the ratio of the two-time matrix
element to the diagonal element reduces to
hΨ(τ1)φ∗(k)φ(k)Ψ(τ1)i2 ≃(1 +
hΨ(τ2)φ∗(k)φ(k)Ψ(τ1)i2
1
k (cid:19)2
4(cid:18) H
(cid:2)eHt1 − eHt2(cid:3)2)−3/2
.
(9)
Evidently, Eq.
(9) is quickly approaching zero which implies that the decoherence of the
system for all t1− t2 ≥ 1/H, once ke−Ht1 and ke−Ht2 are less than H. Thus, the relationship
between the k-mode fluctuations at times t1 and t2 can be treated as classical perturbations
provided that both physical wavelengths, eHt1 /k and eHt2 /k, are super-Hubble sized with the
time separation t1 − t2 being greater than one Hubble time.
The quantum decoherence does not exclude the classical time correlation of perturbations
[7]. The classical correlation can be derived by means of the Wigner distribution functions of
4
EUROPHYSICS LETTERS
the system. For the k mode, it reads
W (φ(k), π(k), τ ) = Z d(cid:16) ϕ
= Nkσk2
π
2π(cid:17) e−iπϕΨ∗(cid:16)φ(k) −
exp(cid:20)− φ(k)2
, τ(cid:17) Ψ(cid:16)φ(k) +
σk(τ )2(cid:21) exp"−σk(τ )2(cid:12)(cid:12)(cid:12)(cid:12)
2# .(10)
Since σk(τ ) is τ -independent when t ≫ 1/H, the first exponent on the right hand side of
Eq.(10) represents the Gaussian probability distribution of the fluctuation of the scalar field
φ(k).
ϕ
2
ϕ
2
, τ(cid:17)
σk(τ )2 φ(k)(cid:12)(cid:12)(cid:12)(cid:12)
π(k) −
F (k, τ )
The time-dependent behavior of the classical perturbations can be extracted through the
second Gaussian probability distribution in Eq. (10). In the classical limit ¯h −→ 0, it yields
W (φ(k), π(k), τ ) ≃ Nk exp(cid:20)− φ(k)2
σk(τ )2(cid:21) δ(cid:18)π(k) −
F (k, τ )
σk(τ )2 φ(k)(cid:19) .
This shows that the classical trajectory of the k mode in phase space is given by
π(k) −
F (k, τ )
σk(τ )2 φ(k) = 0.
(11)
(12)
The degree, or the effectiveness, of classical correlation can be measured by the relative sharp-
ness of the classical trajectory in phase space, which is defined as the ratio of the dispersion
in momentum to the magnitude of the average of the momentum π(k) [8]. By virtue of Eq.
(10), this ratio is
1/σk(τ )
(F (k, τ )/σk(τ )2)φ(k) ≃
1
F (k, τ ) ≪ 1,
if kτ = (k/H) exp(−Ht) ≪ 1.
(13)
Therefore, the classical perturbations of super-Hubble size, i.e. ke−Ht ≪ H, are perfectly
coherent. Namely, after the quantum-to-classical transition, the evolution of the initial per-
turbations can be calculated by the classical trajectory Eq. (12).
The scale tc of the time-dependent correlation between classical perturbations of k mode
can be characterized by
1
tc ≃ −
∂φ(k)/∂t
φ(k)
∂τ
∂t
π(k)
a2φ(k) ≃
= −
k2e−2Ht
H
.
Hence, for super-Hubble sized perturbations, we have
tcH ≫ 1.
(14)
(15)
That is, the correlation time of the primeval perturbations is much longer than the Hubble
time H −1.
This result is anticipated because the only interaction of the inflation in the "standard"
model during inflation is the gravitational particle creation. After the quantum-to-classical
transition, the perturbed scalar field consists of classical free motion. Since there is no cou-
pling between either comoving modes k or physical modes k
p, the evolution of the classical
perturbations of the φ field are coherent. Therefore, one may conclude that the large cor-
relation time of the initial perturbations probably is a generic feature for models with free
inflaton, i.e., besides the gravitational coupling of the inflaton, there are no interactions and
W.-L. Lee and L.-Z. Fang: Time dependent correlations of inflationary perturbations5
self-interaction during the slow-roll phase. This point can be seen more clearly if we establish
an equation for the classical perturbations of the φ field.
Actually, one can regard Eq.(14) as an evolution equation of the ensemble averaged φ(k),
i.e. ∂φ(k)/∂t ≃ −(k2e−2Ht/H)φ(k). The uncertainty of this equation can be estimated by
the time scale given by
1
t′
c ≃(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
∆∂φ(k)/∂t
φ(k)
= ∆π(k)
a3 φ(k) ≃
k3e−3Ht
H 2
.
(16)
We have t′
in phase space is reasonable in average.
c ≫ tc and therefore, the classical evolution equation for the k mode perturbations
The two equations (14) and (16) of the ensemble averaged mode φ(k) can be combined
into the following Langevin-like equation for stochastic variable φ(k)
∂φ(k)
∂t ≃ −
k2e−2Ht
H
φ(k) + ηk,
(17)
where the small Gaussian noise ηk has zero mean with variance ≃ k3/2e−3Ht/H. Therefore,
the ensemble average of (17) yields Eq.(14), and the variance gives (16). Since Eq.(17) is linear
to φ(k, t), it is straightforward to get the equation of φ(r, t) in the comoving r-representation.
After employing the physical coordinates in which dxi = eHtdri (i = 1, 2, 3), Eq.(17) becomes
∂φ(x, t)
∂t
≃
1
H ∇2
xφ(x, t) + η(x, t),
(18)
where ∇2
x stands for the Laplacian of the physical coordinates x. Because the variance of
the stochastic noise is scaled as hη(x, t)2i ∝ e−3Ht, the noise term is strongly suppressed as
Ht ≫ 1.
Equation (18) shows that the correlation time for a mode with the physical wavevector kp,
i.e. φ(x, t) ∼ e−ikp·x, is proportional to k−2
p . This is typical to the diffusion-like soft modes in
the theory for non-equilibrium systems [9]. If a perturbed field consists of the superposition of
such soft long wavelength perturbations described by Eq.(18), the time dependent correlation
function of the perturbed φ field possesses a long-tail [10]. This behavior is more transparent
when considering modes characterized by e−iωt−ikp·x. For these modes Eq.(18) gives rise to
the dispersion relation as ω = −ik2
p → ∞ when
kp → 0, and these modes lead to long-range correlations in general.
Since we are only interested in the perturbations consisting of modes with physical wave-
length larger than H −1, the field φ(x, t) crossing over the horizon can be prescribed by
p/H. Apparently, the relaxation time −1/ℑωk
φ(x, t) =Zkp<H
d3kp φkp e−iωt−ikp·x,
(19)
where φkp is the amplitude of the perturbation for modes kp. Thus, the time-dependent
correlation function is determined by
hφ(x, t1)φ∗(x, t2)i =Zkp<H
d3kph φkp
φ∗
kpi exp −
k2
p
H t1 − t2! .
(20)
At a given instant, the distribution of h φ φ∗i with respect to the kp modes should be the same
as that along the k modes, since the relation between kp and k is specified by a constant
6
EUROPHYSICS LETTERS
φ∗
kpi ∝ k−3
multiplier. On the other hand, Eqs.(4) and (11) imply that, hφ(k)2i ≃ σk(τ )2 ∝ k−3 in the
super-horizon region. Therefore, h φkp
p , and the integral (20) is infrared divergent.
However, the duration of the inflation spans only a finite period of time, there must be an
φ∗
infrared cutoff at k ≃ H/R with R ≫ 1. Thus, h φkp
kpi can be approximated as a quantity
independent of kp, and Eq.(20) yields
t1 − t2(cid:19)3/2
hφ(x, t1)φ∗(x, t2)i ∝(cid:18) tc
t1 − t2 > tc,
,
if
(21)
the correlation time
where the correlation time tc ≃ C/H, and the constant C > 1, i.e.
is longer than the Hubble time. Hence, during the super-Hubble evolution, the field φ is
coherent regardless of the horizon-crossing moments of the perturbations provided that time
difference is less than C/H. This result illuminates that the gravitational particle creation can
be considered as a dissipation source for the non-equilibrium inflaton φ during the slow-roll.
The relaxation of that dissipative process is dominated by soft modes and subsequently gives
rise to the classical perturbations with a long-tailed time correlation. On the other hand, the
soft modes do not induce any long range spatial correlation and therefore, the perturbations in
different spaces x retain themselves. That is, the random field of the inflationary perturbations
is "double-faced": its spatial distribution is Gaussian, while the temporal or scaled distribution
is coherent. This property leads to a measurable effect if we use field variables based on a
space-scale decomposition.
Let us consider such space-scale decomposition given by the bases Ψkp,x(x′) which are
localized, orthogonal and complete. The indices kp and x denote a cell in the phase (x, kp)
space, i.e. kp → kp + ∆kp, x → x + ∆x, and the differential volume element ∆x∆kp ≃ 2π.
Wavelet analysis provides various bases for this sort of space-scale decomposition [11]. With
a suitable wavelet decomposition, the φ field can be described by the variables δφkp,x defined
as
Obviously, δφkp,x represents the amplitude of the perturbation of mode (kp, x), i.e. the fluc-
tuations of the field at position around x and on scale around kp.
According to Eq.(1), the amplitude δφkp,x is mainly attributed to φ(x) with the horizon-
crossing time corresponding to kp. Consequently, the time-time correlation of Eq.(21) implies
hδφkp1,xδφkp2,xi 6= 0,
(23)
if the difference between the horizon-crossing times corresponding to kp1 and kp2 is less than
C/H. Equation (23) is a local (at spatial area around x) scale-scale (kp1 and kp2) correlation.
By means of the variables δφkp,x one can easily perceive this "double-faces" feature of the
inflationary perturbations. The variables δφkp,x which are Gaussian with respect to x entail
hδφkp,xδφkp,x′i = P (kp)δx,x′,
(24)
and all higher order cumulants of δφkp,x are zero, in which P (kp) is the power spectrum of the
Gaussian fluctuations δφkp,x [12]. Meanwhile, the variables at different scales kp may have
carried some correlation. As a simple example, let us consider
δφk′
p,x = αδφkp,x,
(25)
where α is a constant. In this case, the perturbation δφk′
i.e. its variance is P (k′
p) = α2P (kp) and all higher order cumulants of δφk′
p,x with respect to x is also Gaussian,
p,x are zero. However,
δφkp,x =Z φ(x′)Ψkp,x(x′)dx′.
(22)
W.-L. Lee and L.-Z. Fang: Time dependent correlations of inflationary perturbations7
p-kp) correlation at the place is significant, i.e. hδφk′
the scale-scale (k′
p,xδφkp,xi ∝ α. Therefore,
a Gaussian power spectrum P (kp) can coexist with a local scale-scale correlation [Eq.(23)].
It should be emphasized that as a statistical measure, the local scale-scale correlation is
independent of the Gaussian power spectrum. One cannot determine whether there is and/or
how strong it is the local scale-scale correlation directly by their power spectrum.
Recently, the search for the local scale-scale correlation of large scale structure samples has
attracted attention. Using the wavelet technique, the local scale-scale correlations of QSO Lyα
forests [13] and COBE data [14] have been studied. On the other hand, the time-dependent
correlation of the primordial perturbations is sensitive to the interaction of the inflationary
field. It is also sensitive to dissipation during the inflation. For instance, if the cosmic inflation
is thermally dissipated [15], the correlation time of the initial cosmic perturbations will be
significantly changed by a thermal-dissipative term in Eq.(17). Hence, the local scale-scale
correlation behavior of the CMB fluctuations and other relevant samples would be useful to
discriminate among models of inflation.
∗ ∗ ∗
W.-L. L. thanks J.-S. Tsay and S.-Y. Lin for helpful discussion. He also acknowledges
support from the Republic of China National Science Council via Grants No. NSC89-2112-
M-001-060.
REFERENCES
[1] Hawking S., Phys. Lett. B, 115 (1982) 295; Starobinsky A. A., Phys. Lett. B, 117 (1982) 175;
Guth A. H. and Pi S.-Y., Phys. Rev. Lett., 49 (1982) 1110; Grishchuk L. P. and Sidorov Y.
V., Phys. Rev. D, 42 (1990) 3413.
[2] Sakagami M., Prog. Theor. Phys., 79 (1988) 442; Brandenberger R., Laflamme R. and
Mijic M., Mod. Phys. Lett. A, 5 (1990) 2311; Nambu Y., Phys. Lett. B, 276 (1992) 11.
[3] See, e.g. Peebles P. J. E. and Vilenkin A., Phys. Rev. D, 59 (1999) 063505.
[4] See, e.g. Lyth D. H. and Riotto A., Phys. Rep., 314 (1999) 1.
[5] See, e.g. Rey S.-J., Nucl. Phys. B, 284 (1987) 706; Sasaki M., Nambu Y. and Nakao K.,
Nucl. Phys. B, 308 (1988) 868.
[6] Brandenberger R., Nucl. Phys. B, 245 (1984) 328; Polarski D. and Starobinsky A. A.,
Class. Quantum. Grav., 13 (1996) 377.
[7] Halliwell J., Phys. Rev. D, 39 (1989) 2912; Habib S. and Laflamme R., Phys. Rev. D, 42
(1990) 4056.
[8] Morikawa M, Phys. Rev. D, 42 (1990) 2929.
[9] See, e.g. Kirkpatrick T. R. and Belitz D., J. Stat. Phys., 87 (1997) 1307; Goldenfeld N.,
Lectures on Phase Transitions and the Renormalization Group (Addison Wesley, Reading MS)
1992.
[10] Ernst M. H., Hauge E. H.& Van Leeuwen J. M. J., Phys. Rev. Lett., 25 (1970) 1254.
[11] Fang L.-Z. and Thews R., Wavelet in Physics (World Scientific, Singapore) 1998.
[12] Pando J. and Fang L.-Z., Phys. Rev. E, 57 (1998) 3593.
[13] Feng L.-L. and Fang L.-Z., Astrophys. J., 535 (2000) 519; Pando J., Lipa P., Greiner M.
and Fang L.-Z., Astrophys. J., 496 (1998) 9;
[14] Pando J., Valls -- Gabaud D. and Fang L.-Z., Phys. Rev. Lett., 81 (1998) 4568; Mukherjee
P., Hobson M. P. and Lasenby A. N., Mon. Not. R. Astron. Soc., 318 (2000) 1157.
[15] Lee W.-L. and Fang L.-Z., Phys. Rev. D, 59 (1999) 083503; Berera A. and Fang L.-Z.,
Phys. Rev. Lett., 74 (1995) 1912; Lee W.-L. and Fang L.-Z., Int. J. Mod. Phys. D, 6 (1997)
305; Lee W.-L. and Fang L.-Z., Class. Quantum Grav., 17 (2000) 4467.
|
astro-ph/0607622 | 1 | 0607 | 2006-07-27T18:27:31 | Formation and Collapse of Nonaxisymmetric Protostellar Cores in Planar Magnetic Interstellar Clouds: Formulation of the Problem and Linear Analysis | [
"astro-ph"
] | We formulate the problem of the formation and collapse of nonaxisymmetric protostellar cores in weakly ionized, self-gravitating, magnetic molecular clouds. In our formulation, molecular clouds are approximated as isothermal, thin (but with finite thickness) sheets. We present the governing dynamical equations for the multifluid system of neutral gas and ions, including ambipolar diffusion, and also a self-consistent treatment of thermal pressure, gravitational, and magnetic (pressure and tension) forces. The dimensionless free parameters characterizing model clouds are discussed. The response of cloud models to linear perturbations is also examined, with particular emphasis on length and time scales for the growth of gravitational instability in magnetically subcritical and supercritical clouds. We investigate their dependence on a cloud's initial mass-to-magnetic-flux ratio (normalized to the critical value for collapse), the dimensionless initial neutral-ion collision time, and also the relative external pressure exerted on a model cloud. Among our results, we find that nearly-critical model clouds have significantly larger characteristic instability lengthscales than do more distinctly sub- or supercritical models. Another result is that the effect of a greater external pressure is to reduce the critical lengthscale for instability. Numerical simulations showing the evolution of model clouds during the linear regime of evolution are also presented, and compared to the results of the dispersion analysis. They are found to be in agreement with the dispersion results, and confirm the dependence of the characteristic length and time scales on parameters such as the initial mass-to-flux ratio and relative external pressure. | astro-ph | astro-ph |
Formation and Collapse of Nonaxisymmetric Protostellar Cores in Planar
Magnetic Interstellar Clouds: Formulation of the Problem and Linear Analysis
To appear in The Astrophysical Journal
Glenn E. Ciolek1,2 and Shantanu Basu3
ABSTRACT
We formulate the problem of the formation and collapse of nonaxisymmetric pro-
In our
tostellar cores in weakly ionized, self-gravitating, magnetic molecular clouds.
formulation, molecular clouds are approximated as isothermal, thin (but with finite
thickness) sheets. We present the governing dynamical equations for the multifluid
system of neutral gas and ions, including ambipolar diffusion, and also a self-consistent
treatment of thermal pressure, gravitational, and magnetic (pressure and tension) forces.
The dimensionless free parameters characterizing model clouds are discussed. The re-
sponse of cloud models to linear perturbations is also examined, with particular empha-
sis on length and time scales for the growth of gravitational instability in magnetically
subcritical and supercritical clouds. We investigate their dependence on a cloud's ini-
tial mass-to-magnetic-flux ratio µ0 (normalized to the critical value for collapse), the
dimensionless initial neutral-ion collision time τni,0, and also the relative external pres-
sure exerted on a model cloud Pext. Among our results, we find that nearly-critical
model clouds have significantly larger characteristic instability lengthscales than do
more distinctly sub- or supercritical models. Another result is that the effect of a
greater external pressure is to reduce the critical lengthscale for instability. Numerical
simulations showing the evolution of model clouds during the linear regime of evolution
are also presented, and compared to the results of the dispersion analysis. They are
found to be in agreement with the dispersion results, and confirm the dependence of
the characteristic length and time scales on parameters such as µ0 and Pext.
Subject headings: diffusion -- ISM: clouds -- ISM: magnetic fields -- MHD -- stars:
formation
1New York Center for Studies on the Origin of Life (NSCORT).
2Department of Physics, Applied Physics, and Astronomy, Rensselaer Polytechnic Institute, 110 8th Street, Troy,
NY 12180; [email protected].
3Department of Physics and Astronomy, University of Western Ontario, London, Ontario N6A 3K7, Canada;
[email protected].
-- 2 --
1.
Introduction
1.1. Background
In recent years there has been a steady and significant accumulation of observational data that
support the idea that magnetic fields play a pivotal role in the evolution of interstellar molecular
clouds and star formation. For instance, polarization maps of extended regions in clouds reveal
large-scale magnetic fields that in several instances are aligned with the short axis of clouds (e.g.
Pereyra & Magalhaes 2004), which would be consistent with support from magnetic forces perpen-
dicular to field lines. Infrared, far-infrared, and submillimeter maps of more localized regions show
ordered field configurations with curvature, including "hourglass" shapes, that suggest dynamic
interaction of magnetic fields and molecular cloud gas (Schleuning 1998; Schleuning et al. 2000;
Fujiyoshi et al. 2001).
Interferometric polarization maps of molecular cloud cores by Lai et al. (2001, 2002) indicate
similar ordered field structures. Furthermore, using a modification of the Chandrasekhar-Fermi
(CF) method (Chandrasekhar & Fermi 1953), they estimated magnetic field strengths and ener-
gies that were large enough to exceed observed gas turbulence in their survey targets. A modified
CF method was applied to the sub-mm polarization measurements of several prestellar cores by
Crutcher et al. (2004), and they found that the mass-to-magnetic flux ratios of their cores were
clustered about (within a factor ∼ 2 above and below) the critical value for gravitational collapse.
This result is consistent with earlier OH Zeeman determinations of the mass-to-flux ratios of proto-
stellar cores (Crutcher 2004). Likewise, Curran et al. (2004) obtained sub-mm polarimetric images
of two high-mass protostellar cores; a CF analysis of their data yielded mass-to-flux ratios for their
two objects that were also nearly equal to the critical value for collapse. Finally, comparing CF
results with OH Zeeman measurements for the immediate environment of the L184 prestellar core,
Crutcher (2004) concludes that it may be an example of a nearly or barely critical core contained
within a magnetically subcritical (i.e., supported) envelope. The grand picture painted by the vari-
ous observational data above is one in which magnetic fields to a large degree control the structure
and dynamics of star-forming molecular clouds, and has many elements of the original, unified
theory of magnetically regulated star formation put forth early on by Mouschovias (1976, 1977,
1978, 1979; see, also, the review by Mouschovias & Ciolek 1999).
Ambipolar diffusion -- the drift of neutral matter with respect to plasma and magnetic fields
-- initiates protostellar core formation in magnetically supported clouds. That is, because of im-
perfect collisional coupling between neutrals and charged particles (including dust grains; Ciolek &
Mouschovias 1993, 1994), gravitationally-driven diffusion of matter in the inner flux tubes of clouds
redistributes mass and magnetic flux that leads to the formation and subsequent dynamical collapse
of supercritical cores or fragments within massive, subcritical envelopes. This process was exhib-
ited in the two-dimensional axisymmetric magnetic cloud simulations of Fiedler & Mouschovias
(1993), and the axisymmetric thin-disk model calculations by Ciolek & Mouschovias (1994, 1995)
and Basu & Mouschovias (1994, 1995a,b). Uniformly, these studies demonstrated the formation of
-- 3 --
supercritical protostellar cores embedded in subcritical envelopes. They were also able to follow
the resultant collapse of these cores well into the stage of dynamical infall, ending at the time when
the central density was enhanced by a factor ∼ 106.
Ciolek & Basu (2000; hereafter CB00) applied one of their axisymmetric magnetic disk models
of core formation by ambipolar diffusion to the L1544 starless core. This model provided a good
theoretical fit to the extended, subsonic infall profile as well as the distribution of mass in that core
found by the earlier, complementary studies of Tafalla et al. (1998) and Williams et al. (1999).
The predicted magnetic field strength in the CB00 model was later found to be in agreement with
OH Zeeman observations of L1544 (Crutcher & Troland 2000; CB00). An analysis of submillimeter
and millimeter maps and theoretical models of the dust temperature distribution in L1544 were
found to be consistent with the emission distribution calculated from the CB00 model (Zucconi,
Walmsley, & Galli 2001). Subsequent detailed, multispectral line maps used to further examine the
kinematics within the L1544 core have also shown reasonable agreement between measured infall
speeds and the CB00 predictions (Caselli et al. 2002). Finally, a multispectral survey by Crapsi
et al. (2004) has found that some of the infall features of the prestellar core L1521F (which, like
L1544, is in the Taurus complex) are comparable to those of the CB00 model.
However, despite the apparent successes of the aforementioned ambipolar diffusion models in
describing the earliest stages of protostellar collapse, they are inherently limited because of their
underlying assumption of axisymmetry. Axisymmetry is clearly a highly restrictive and idealized
assumption, and is not likely to naturally occur in molecular clouds. In fact, sub-mm maps of star-
forming clouds frequently indicate distinctively inhomogeneous and irregular large-scale structure
and multiple cores (e.g., Andr´e, Motte, & Belloche 2001; Motte, Andr´e, & Neri 1998). Investigations
at these wavelengths reveal irregularities on the scale of individual cores as well (e.g., Bacmann et
al. 2000). Moreover, the observational studies of the L1544 core cited above find that it is definitely
not axisymmetric. This is not a surprising result, as statistical analyses of data sets and catalogs
of cores (e.g., Jones, Basu, & Dubinski 2001; Jones & Basu 2002; Kerton et al. 2003) reveal that
their shapes can generally be best fit with a distribution of triaxial ellipsoids, which are typically
more oblate than prolate.
Indebetouw & Zweibel (2000) simulated the formation of supercritical cores by ambipolar
diffusion in infinitesimally thin, magnetic layers. Their models were two-dimensional and nonax-
isymmetric, and they focused primarily on clouds that were initially subcritical. They followed the
evolution of cores to when the maximum column density was a factor . 5 above the initial mean
column density, and the irregular cores that developed were reminiscent of the sub-mm observa-
tions cited above.1 They also showed that the cores in their calculations developed on a timescale
1Some of the results in the interiors of their cores (scaling of magnetic field strength with central density, subsonic
infall speeds, etc.) were similar to those found in the earlier axisymmetric model simulations over the same range
of central density. This is also true for the nonaxisymmetric critical cloud model of Basu & Ciolek (2004 -- see the
denser, inner core regions in their Fig. 1). This agreement is probably the reason why the axisymmetric CB00 model
-- 4 --
corresponding to the maximum linear growth rate of gravitational instability, and had a size equal
to the corresponding wavelength (Zweibel 1998; see, also, § 3 below). Basu & Ciolek (2004; here-
after BC04) presented numerical models displaying the gravitational collapse of multiple dense,
asymmetric protostellar cores in thin (yet, not infinitesimally so) planar magnetic clouds. One
of their model clouds had an initial reference state with a mass-to-magnetic flux ratio that was
exactly the critical value (µ0 = 1, see §§ 2.2 and 3.1 below). Another model of theirs was initially
supercritical by a factor of 2 (i.e., µ0 = 2). They found that, in the initially critical model, transfer
of mass by ambipolar diffusion resulted in the eventual formation of supercritical cores contained
within subcritical envelopes, similar to that which occurred in their earlier axisymmetric models
that started with initially subcritical conditions. By contrast, the initially supercritical model of
BC04 had a much more rapid and dynamical infall, with significantly greater infall speeds (sonic
and supersonic) extending over much larger regions. Based on the physically distinct and different
predictions of these two models, they noted that molecular clouds that are supercritical by more
than a factor ∼ 2 would be incompatible with the generally observed subsonic infall motions.
1.2. Outline
In this paper, we present the formulation for modeling the formation and nonaxisymmetric
collapse of protostellar cores in planar magnetic clouds. In § 2 we describe the fundamental assump-
tions and derive the necessary system of governing equations for a model cloud. The equations are
put in nondimensional form, and the resulting free parameters of a model and their physical mean-
ing are described. Our numerical method of solving the governing equations is also discussed. To
provide a basis for understanding the underlying physics of ambipolar diffusion and gravitational
collapse in magnetic sheets, the stability of model clouds is examined in § 3 by linearizing and
Fourier-analyzing the governing equations. The results of the stability analysis are also compared
with full numerical simulations of models in the limit of small-amplitude perturbations. In section
§ 4 we summarize our results, and discuss their relevance to forthcoming fully nonlinear studies of
nonaxisymmetric core formation and collapse.
2. Physical Formulation
We model clouds as isothermal thin sheets with temperature T , embedded in a hot and tenuous
external medium of constant pressure Pext. This simplifying assumption, the thin sheet approxima-
tion, is based on the numerical simulations of ambipolar-diffusion-initiated formation of cores by
Fiedler & Mouschovias (1993), who found that two-dimensional, axially symmetric magnetically
supported model molecular clouds rapidly flatten and establish force-balance along the direction
of magnetic field lines, even while contraction and dynamical collapse took place in the direction
perpendicular to the field. These results were soon afterward incorporated in the protostellar core
is a reasonable physical model for the interior of the asymmetric L1544 prestellar core.
-- 5 --
formation studies of Ciolek & Mouschovias (1993; hereafter CM93) and Basu & Mouschovias (1994;
hereafter BM94), who modeled molecular clouds as thin axisymmetric disks threaded by a vertical
magnetic field, with hydrostatic equilibrium maintained along field lines at all times. Here we again
adopt the formulation of CM93 and BM94, but now forego the restriction of axisymmetry.
The appealing feature of the thin-sheet approximation is that it can be used to model the
physical processes necessary to core and star formation in a theoretically tractable and realistic
way, including a self-consistent treatment of both magnetic pressure and tension forces. Another
advantage of thin-sheet models is that because the physical equations have been simplified by
integration in one dimension (here, in the direction of the vertical magnetic field), the problem has
been substantially reduced from its full three-dimensional complexity. Because of this, sheet models
can provide significantly higher spatial resolution, and require much less computational time and
storage than fully three-dimensional cloud models. There is also some observational justification for
the use of the thin sheet approximation. The often observed alignment, or at least close correlation,
between the projected magnetic field direction from polarization measurements and the core minor
axis is evidence for flattening along the field (Basu 2000). Furthermore, an analysis of core shapes
by Jones et al. (2001) and Jones & Basu (2002) implies that they are preferentially flattened along
one direction.
Of course, sheet models may not account for all of the observed morphological features of
molecular clouds and their envelopes. Especially for those with large internal velocity dispersion,
which can provide substantial support against self-gravity, and therefore be extended along magnetic
field lines. It turns out though, that even in clouds or complexes with such substantial velocity
dispersion or turbulence, thin-sheet models may still be used, so long as the application is restricted
to dense sub-regions of the cloud. This has been demonstrated by Kudoh & Basu (2003, 2006),
who analyzed the nonlinear support of stratified molecular clouds due to an ensemble of driven
hydromagnetic waves. They found that because of density stratification, the largest, supersonic,
velocities are in the low density envelope of a cloud, while a dense region near the midplane has
transonic or subsonic motions. Hence, in this situation, thin-sheet models may reasonably be
applied to a high-density fragment or sub-cloud, which will most likely be the site of subsequent
star formation.
The most significant systematic source of possible inaccuracy of a thin-sheet model is that it
typically overestimates the strength of the gravitational field in the equatorial plane (i.e., the plane
of the disk) when compared to less condensed or concentrated systems. For instance, the critical
wavelength for gravitational instability in very thin non-magnetic clouds is found to be half the
value of the critical length found for an equilibrium layer with an exponential atmosphere. This can
be seen by comparing our equation (40) derived in § 3.1, in the limit of zero external pressure, to
equations (13-36) and (13-42) of Spitzer (1978). However, the equilibrium layer calculation assumes
an infinite vertical extent, which reduces the system's total gravitational binding energy; this is
therefore an upper limit to the possible overestimate of calculating the gravitational field by using
a thin disk. A better estimate was made by Kiguchi et al. (1987, see their Fig. 11), who showed
-- 6 --
that the magnitude of the planar gravitational field for an infinitesimally thin disk exceeds that
of a flattened cloud of finite extent by at most only ∼ 40%, with the largest discrepancy between
the two occurring for a very small region around the cloud center. BM94 presented a method of
correcting for finite-thickness effects in the magnetic thin-disk approximation, deriving a quadratic
correction (in terms of the local disk thickness) to the gravitational field. For the density regimes
we intend to study (. 107 cm−3), the BM94 correction had the effect of altering the evolution of
a model's core by about 10% during the early stages of core evolution, to . 30% at much later
stages. Finally, we note that because the planar magnetic field in the sheet is calculated the same
way that the planar gravitational field is (see eqs. [15] - [16b] and [20] - [21b] below), there is also
a comparable overestimate in the magnitude of the magnetic tension force. Since the magnetic
tension force opposes the gravitational force, the overestimate in the value of the gravitational
field due to the thin-sheet approximation is offset by a corresponding overestimate of the planar
magnetic field, thereby reducing the net effect of overestimating both fields.
Thin-sheet models have been used by many workers in the field of molecular clouds and pro-
tostars. They were used early on by Narita, Hayashi, & Miyama (1984) to study star formation
in axisymmetric, non-magnetic clouds. As mentioned above, the thin-sheet approximation was
developed and applied to ambipolar diffusion and core formation in isothermal magnetic molecular
clouds by CM93 (who included the effects of dust) and BM94 (who included rotation and mag-
netic braking). It was also later adopted by Li & Shu (1997a,b), who studied the equilibria and
self-similar gravitational collapse of clouds with frozen-in magnetic flux (i.e., no diffusion). The
approach to self-similar collapse during the later stage of core collapse with ambipolar diffusion
in thin-disk clouds was described by Basu (1997). Ciolek & Konigl (1998) incorporated the ef-
fect of the formation of a central gravitating point mass (a protostar) in numerical simulations of
ambipolar-diffusion-driven dynamical core collapse in axisymmetric thin-disk clouds; a self-similar
solution to this same problem was also provided by Contopoulos, Ciolek, & Konigl (1998). Tassis
& Mouschovias (2005a,b) also investigated this particular topic, further extending the analysis by
including a detailed calculation of multi-fluid effects on the conductivity of the infalling gas during
the later stages of supercritical core collapse and accretion.
A nonaxisymmetric 'toy model' idealization of a magnetically critical, flux-frozen turbulent
sheet-like molecular cloud was suggested by Allen & Shu (2000). As noted in §1.1, Indebetouw &
Zweibel (2000) first presented nonaxisymmetric simulations of core formation by ambipolar diffusion
in infinitesimally thin magnetic sheets. BC04 presented models of nonaxisymmetric, gravitationally
collapsing cores in magnetically critical and supercritical finite-thickness sheet-like molecular clouds,
including ambipolar diffusion and its consequent effect on the dynamical evolution of clouds and
cores. Li & Nakamura (2004) and Nakamura & Li (2005) used the thin-sheet approximation to
examine the combined effect of turbulent initial conditions and ambipolar diffusion in forming
supercritical cores.
-- 7 --
2.1. Fundamental Equations
We present the necessary system of equations to model core formation in weakly ionized,
magnetic interstellar clouds. As stated above, our model clouds are taken to be thin, with local
vertical half-thickness Z(x, y, t) in a Cartesian coordinate system. By a sheet being thin we mean
that for any physical quantity f (x, y, z, t), the criterion f /∇pf ≫ Z is always satisfied, where
∇p ≡ x∂/∂x + y∂/∂y is the planar gradient operator. The magnetic field threading a cloud is
taken to have the form
B(x, y, z, t) = Bz,eq(x, y, t)z
for z ≤ Z(x, y, t),
B(x, y, z, t) = Bz(x, y, z, t)z + Bx(x, y, z, t)x + By(x, y, z, t)y
for z > Z(x, y, t),
(1a)
(1b)
where Bz,eq is the magnetic field strength in the equatorial plane of the cloud. In the limit z → ∞,
B → Bref z, where Bref is a constant, uniform reference magnetic field very far away from the sheet.
From now on, all physical quantities are understood to be a function of time t.
Note that the condition on the planar gradient of physical quantities within the thin-sheet
approximation implies that Z is the lower limit to the scales that can be described by our model.
(For the gravitationally unstable modes, this condition is always satisfied, as the critical lengths
always exceed this value -- see eq. [40].)
The unit normal vector to the upper and lower surfaces of the sheet is
n(x, y, ±Z) =
±z ∓ [(∂Z/∂x)x + (∂Z/∂y)y]
[1 + (∂Z/∂x)2 + (∂Z/∂y)2]1/2
.
(2)
Use of the integral form of Gauss's law yields the continuity equation for the normal component of
the magnetic field across the upper and lower surfaces of the sheet,
Bz(x, y, ±Z) − Bx(x, y, ±Z)
∂Z
∂x
− By(x, y, ±Z)
∂Z
∂y
= Bz,eq(x, y).
(3)
The system of multifluid equations necessary to model a weakly-ionized, magnetic, self-gravitating,
isothermal molecular cloud is given by equations (9a)-(9m) of CM93. (In this study we ignore the
dynamical effect of interstellar dust grains. Hence, those terms representing grain contributions
appearing in the system of fluid equations in CM93 are neglected.) As was done in CM93 and
BM94, we simplify these basic equations in the thin-sheet approximation by vertically integrating
them from z = −Z(x, y) to z = +Z(x, y). Doing so for the equation of mass continuity yields
∂σn
∂t
+
∂
∂x
(σnvn,x) +
∂
∂y
(σnvn,y) = 0 ,
(4)
where σn(x, y) ≡R +Z
−Z ρn(x, y)dz is the mass column density, and vn,x and vn,y are respectively the
x- and y-components of the neutral velocity. In deriving equation (4) we have used the chain rule
to obtain the velocity of the surface of the disk,
dZ
dt
= vn,x
∂Z
∂x
+ vn,y
∂Z
∂y
.
(5)
-- 8 --
We have also employed a "one-zone" approximation, where we assume z-independence of physical
quantities such as ρn, and the planar velocity components vn,x, and vn,y. This approximation is
also used for the planar components of the gravitational acceleration gx and gy that appear in the
equations below.
To derive the x- and y-components of the force equation (per unit area) for the neutrals, we
use equation (2) and the total (thermal plus Maxwell) stress tensor
T = −(cid:18)ρnC 2 +
B2
8π(cid:19) 1 +
BB
4π
,
(6)
where 1 is the identity tensor, C = (kBT /mn)1/2 is the isothermal speed of sound; kB is the
Boltzmann constant, and mn is the mean mass of a neutral particle (= 2.33 a.m.u. for an H2 gas
with a 10% He abundance by number). Using equation (3), the symmetry conditions on Bz,eq and
Bz and the antisymmetry conditions on Bx and By about the equatorial plane, along with the
divergence theorem, the integrated force equations are
where
Fmag,x =
Fmag,y =
∂
∂t
∂
∂t
(σnvn,x) +
(σnvn,y) +
∂
∂x
∂
∂x(cid:0)σnv2
n,x(cid:1) +
(σnvn,yvn,x) +
∂
∂y
(σnvn,xvn,y) = σngx − C 2
eff
∂
∂y (cid:0)σnv2
n,y(cid:1) = σngy − C 2
eff
∂σn
∂x
∂σn
∂y
+ Fmag,x ,
+ Fmag,y ,
(7a)
(7b)
∂Bz,eq
Bz,eq
2π (cid:18)Bx,Z − Z
+
1
4π
∂Z
∂x "B2
x,Z + B2
∂Bz,eq
Bz,eq
2π (cid:18)By,Z − Z
+
1
4π
∂Z
∂y "B2
x,Z + B2
∂x (cid:19)
y,Z + 2Bz,eq(cid:18)Bx,Z
∂y (cid:19)
y,Z + 2Bz,eq(cid:18)Bx,Z
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19) +(cid:18)Bx,Z
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19)2# , (8a)
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19) +(cid:18)Bx,Z
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19)2# , (8b)
Bx,Z ≡ Bx(x, y, +Z), By,Z ≡ By(x, y, +Z), and
C 2
eff ≡
π
2
Gσ2
n(cid:0)3Pext + π
(cid:0)Pext + π
2 Gσ2
2 Gσ2
n(cid:1)
n(cid:1)2 C 2
(9)
is the local effective sound speed. G is the gravitational constant. The second term on the right
hand sides of equations (8a) and (8b) represent small modifications to the magnetic force due to
planar gradients of the half-thickness Z.
The equations for vi,x and vi,y, the x- and y-components of the ion velocity, are similarly
obtained by vertical integration of the force equation for the ions. They are, respectively,
vi,x = vn,x +
vi,y = vn,y +
τni
σn
τni
σn
Fmag,x ,
Fmag,y ,
(10a)
(10b)
where the neutral-ion collision (momentum-exchange) time
-- 9 --
τni = 1.4(cid:18) mi + mH2
mi
(cid:19)
1
nihσwiiH2
.
(11)
The quantity mi is the ion mass, which we take to be 25 a.m.u., the mass of the typical atomic
(Na+, Mg+) and molecular (HCO+) ion species in clouds; hσwiiH2 is the neutral-ion collision rate,
and is equal to 1.69× 10−9 cm3 s−1 for H2 − HCO+ collisions (McDaniel & Mason 1973). The factor
of 1.4 in equation (11) accounts for the fact that the inertia of helium is neglected in calculating the
slowing-down time of the neutrals by collisions with ions. (Further discussion on this point can be
found in § 2.1 of Mouschovias & Ciolek 1999.) For the ion number density we assume a power-law
behavior of the form
ni = K(cid:16)
nn
105 cm−3(cid:17)k
,
(12)
where K (≃ 3×10−3 cm−3) and k (≃ 1/2) are constants. In reality, the exponent k is also a function
of density (Ciolek & Mouschovias 1998), due to the fact that ambipolar diffusion can deplete the
abundance of dust grains in a contracting core (Ciolek & Mouschovias 1996), which alters the rate
of capture and recombination of ions and electrons on grain surfaces. However, we ignore this effect
in our models for the time being.
Integrating the z-component of the force equation for the neutrals from z = 0 to z = +Z, and
requiring that there be hydrostatic equilibrium along field lines yields
ρnC 2 =
π
2
n + Pext + (cid:18)B2
Gσ2
x,Z + B2
y,Z +hBx,Z
8π
∂Z
∂x + By,Z
∂Z
∂yi2(cid:19)
,
(13)
where we have used the Gaussian relation for a thin sheet, gz(x, y, +Z) = −2πGσn(x, y, +Z). As
discussed in CM93 and BM94, the first two terms on the right side of equation (13) represent the
self-gravitational stress and external pressure acting on a sheet, respectively. The latter term --
not included in our earlier studies (e.g., see eq. [26] of CM93) -- is the total magnetic "pinching"
term due to magnetic pressure and tension stresses that act to compress the sheet. Although this
last term is generally smaller than the others, we retain it nevertheless in our models, since it costs
very little computationally to include it.
We use Poisson's equation for a very thin sheet to solve for the gravitational potential ψ:
∇2ψ(x, y, z) = 4πGσn(x, y)δ(z).
(14)
Imposing the boundary condition limz→∞ ψ(x, y, z) → 0, equation (14) can be solved by the
method of Fourier transforms (e.g., Wyld 1976; Byron & Fuller 1992). Doing so, one finds for
z = 0,
F[ψ(x, y, 0)] = −2πG
F[σn(x, y)]
kz
,
(15)
1
where F[f ] is the two-dimensional Fourier transform of the function f , and kz = (k2
2 is
a function of the planar wave numbers kx and ky. Hence, at any time t we can determine F[ψ]
x + k2
y)
-- 10 --
by calculating F[σn]. Inverting the transform then yields ψ at that time, and from that we can
calculate the gravitational field
gx(x, y, 0) = −
∂ψ(x, y, 0)
∂x
,
gy(x, y, 0) = −
∂ψ(x, y, 0)
∂y
.
(16a)
(16b)
The magnetic field components Bx,Z and By,Z can be gotten in a similar fashion. Above the
sheet, the magnetic field can be written in the form B(x, y, z > +Z) = B′(x, y, z) + Bref z, where
B′ is the reduced magnetic field. The assumption that the constant pressure external medium is
hot and tenuous implies that it is also current-free (jext = 0). Consequently, Ampere's law gives
∇ × B = 0 in the region above the sheet. Therefore, B′ = −∇Ψ, where Ψ is a scalar magnetic
potential, that, because of Gauss's law (∇ · B = 0), satisfies Laplace's equation,
∇2Ψ(x, y, z > +Z) = 0.
(17)
From our earlier comments on the external field we have B′ → 0 very far from the sheet. Therefore,
lim
z→∞
Ψ(x, y, z) = 0 .
(18)
The boundary condition on Ψ at z = +Z is derived from the continuity equation (3) for the normal
component of the magnetic field across the top surface of the sheet, which is, written in terms of
Ψ,
∂Ψ(x, y, +Z)
∂z
−
∂Ψ(x, y, +Z)
∂x
∂Z
∂x
−
∂Ψ(x, y, +Z)
∂y
∂Z
∂y
= − [Bz,eq(x, y) − Bref ]
.
(19a)
In the thin-sheet approximation, this becomes
lim
+Z→0
∂Ψ(x, y, +Z)
∂z
= − [Bz,eq(x, y) − Bref]
.
(19b)
The solution of the magnetic Laplace equation (17) for Ψ, subject to the boundary conditions
(18) and (19b), can also be performed by Fourier transform, in analogy to what is done for the
gravitational potential ψ. In the limit +Z → 0, we find
F[Ψ(x, y, 0)] =
F [Bz,eq(x, y) − Bref]
kz
.
Once Ψ is obtained, by inverting the transform, it follows that
Bx,Z = lim
+Z→0
−
By,Z = lim
+Z→0
−
∂Ψ(x, y, +Z)
∂x
∂Ψ(x, y, +Z)
∂y
,
.
(20)
(21a)
(21b)
-- 11 --
To close our system of equations, the evolution of the equatorial magnetic field in the plane
of the sheet Bz,eq(x, y) is governed by the magnetic induction equation. For the density range we
consider in our cloud models, 103 cm−3 . nn . 108 cm−3, the magnetic field is effectively "frozen"
in the ion-electron plasma (see §§ 2.2 - 2.4 of Mouschovias & Ciolek 1999 for a discussion). Hence,
advection of magnetic flux in our system is described by
∂Bz,eq
∂t
= −
∂
∂x
(Bz,eqvi,x) −
∂
∂y
(Bz,eqvi,y)
.
(22)
2.2. Boundary Conditions, Uniform Background State and Initial Conditions
As described in the preceding section, model clouds are assumed to be isothermal thin planar
sheets of infinite extent. We follow the evolution in a square Cartesian region of size L, with
−L/2 ≤ x ≤ L/2, −L/2 ≤ y ≤ L/2. Periodic boundary conditions are used for all physical
quantities.
Model clouds are initially characterized by a static, uniform background state with constant
column density σn,0 and equatorial magnetic field Bref z (i.e., Bz,eq,0 = Bref). From equations
(7a), (7b), (16a), (16b), (21a), and (21b), it is seen that all forces -- thermal, gravitational, and
magnetic -- are identically equal to zero in the background state. Evolution is initiated in a cloud
by superposing a set of random column density perturbations, δσn(x, y) (≪ σn,0, typically) on the
uniform background state at time t = 0. To maintain the same local mass-to-flux ratio σn/Bz,eq
in the initial state as in the reference state, the magnetic field is simultaneously perturbed, with
δBz,eq/Bref = δσn/σn,0.
2.3. Dimensionless Equations and Free Parameters
The actual system of equations we solve are dimensionless versions of the equations presented
in § 2.1 above. We adopt the following normalizations: the velocity unit is [v] = C, the column
density unit is [σ] = σn,0, unit of acceleration is [a] = 2πGσn,0, the time unit is [t] = C/2πGσn,0, and
the length unit is [l] = C 2/2πGσn,0. From this system one can also construct a unit of magnetic
field strength, [B] = 2πG1/2σn,0. With these normalizations, the dimensionless equations that
govern the evolution of a model cloud are
(σnvn,x) −
(σnvn,y) ,
∂σn
∂t
= −
(σnvn,x) = −
(σnvn,y) = −
∂
∂t
∂
∂t
∂
∂x
∂
∂
∂x
∂
∂y
∂
∂y
∂x(cid:0)σnv2
n,x(cid:1) −
(σnvn,yvn,x) −
(σnvn,yvn,x) + σngx − C 2
eff
n,y(cid:1) + σngy − C 2
eff
∂
∂y (cid:0)σnv2
∂x (cid:19)
∂Bz,eq
Fmag,x = Bz,eq(cid:18)Bx,Z − Z
∂σn
∂x
∂σn
∂y
+ Fmag,x ,
+ Fmag,y ,
(23a)
(23b)
(23c)
+
1
2
∂Z
∂x "B2
x,Z + B2
Fmag,y = Bz,eq(cid:18)By,Z − Z
-- 12 --
∂Bz,eq
y,Z + 2Bz,eq(cid:18)Bx,Z
∂y (cid:19)
y,Z + 2Bz,eq(cid:18)Bx,Z
x,Z + B2
+
1
2
∂Z
∂y "B2
n(cid:17)
n(cid:16)3 Pext + σ2
n(cid:17)2
(cid:16) Pext + σ2
,
C 2
eff = σ2
∂Bz,eq
∂t
∂
∂x
= −
(Bz,eqvi,x) −
∂
∂y
(Bz,eqvi,y)
,
τni,0
ρn,0(cid:19)k
σn (cid:18) ρn
ρn,0(cid:19)k
σn (cid:18) ρn
τni,0
Fmag,x ,
Fmag,y
n + Pext + B2
x,Z + B2
y,Z +(cid:20)Bx,Z
vi,x = vn,x +
vi,y = vn,y +
ρn =
Z =
1
4 σ2
σn
2ρn
,
F[ψ] = −
gx = −
gy = −
,
F[σn]
kz
∂ψ
∂x
∂ψ
∂y
,
,
F[Ψ] =
F[Bz,eq − Bref ]
,
Bx,Z = −
By,Z = −
kz
,
,
∂Ψ
∂x
∂Ψ
∂y
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19) +(cid:18)Bx,Z
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19)2# , (23d)
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19) +(cid:18)Bx,Z
∂Z
∂x
+ By,Z
∂Z
∂y(cid:19)2# , (23e)
∂Z
∂x
+ By,Z
∂Z
∂y(cid:21)2! ,
(23f)
(23g)
(23h)
(23i)
(23j)
(23k)
(23l)
(23m)
(23n)
(23o)
(23p)
(23q)
where ρn,0 = (1 + Pext)/4 is the dimensionless neutral mass density of the background state, and ψ
and Ψ represent their values in the equatorial plane.
The above equations are, for the most part, the Cartesian analogs of the axisymmetric thin-
disk equations presented in CM93 (their [66a]-[66q]) and BM94 (their [34a]-[34m]). A similar set
of non-ideal MHD equations were used to model nonaxisymmetric core formation by Indebetouw
& Zweibel (2000), with one particular exception: they modeled clouds as infinitesimally thin, with
Z = 0, and neglected the effect of magnetic pressure (i.e., the terms dependent on Z in the equations
above, and also in CM93 and BM94). This means that the stabilizing effect of magnetosound modes
-- 13 --
were not accounted for in the models of Indebetouw & Zweibel. Although this may be valid in some
cases, there are, as we show below, some regions in the physically allowed parameter space for clouds
in which magnetic pressure cannot be considered negligible compared to magnetic tension. The
omission of magnetic pressure in these instances leads to inaccuracies in the length- and timescales
for the onset of gravitational instability in magnetic clouds.
Equations (23a)-(23q) have several non-dimensional parameters. Pext ≡ 2Pext/πGσ2
n,0 is the
ratio of the external pressure acting on the sheet to the vertical self-gravitational stress of the
background state. The effect of ambipolar diffusion is expressed by the dimensionless initial neutral-
ion collision time, τni,0 ≡ 2πGσn,0τni,0/C. The limit τni,0 → ∞ corresponds to extremely poor
neutral-ion collisional coupling, so that the ions and magnetic field have no effect on the neutrals.
In the opposite limit, τni,0 = 0, the neutrals are perfectly coupled to the ions, due to frequent
collisions, and the magnetic field will be essentially frozen in the neutral matter. The parameter k
(=1/2, typically) is the exponent in the power-law expression (12) used to calculate the ion density
as a function of the neutral density. As we noted earlier, this constant power-law assumption
is only an approximation, since ambipolar diffusion has been shown to make k a function of nn
in models with a more realistic ion chemistry network (Ciolek & Mouschovias 1998). Finally,
Bref = Bref/2πG1/2σn,0 is the dimensionless magnetic field strength of the background state. For
the units we have chosen for the column density and the magnetic field, the dimensionless mass-
to-magnetic-flux ratio of the background state is
µ0 ≡ 2πG
1
2 σn,0
Bref
=
1
Bref
.
(24)
This also happens to be the mass-to-flux ratio in units of the critical value for gravitational collapse,
1
2 (Nakano & Nakamura 1978; see, also, § 3.1 below). Models with µ0 < 1 ( Bref > 1) are
1/2πG
subcritical clouds, while those with µ0 > 1 ( Bref < 1) are supercritical.
Normally, we set σn,0 by specifying the temperature T and the density nn,0 of the background
state. Using equation (13), and the fact that C = 0.188(T /10 K)1/2 km s−1 for mn = 2.33 a.m.u.,
σn,0 =
3.63 × 10−3
1
(1 + Pext)
2 (cid:16)
nn,0
103 cm−3(cid:17)
1
2 (cid:18) T
10 K(cid:19)
1
2
g cm−2 .
(25)
Both C and σn,0 are normalizing units that we have adopted for our cloud models. The units for
length and time have the scalings
and
[l] = 7.48 × 10−2(cid:18) T
10K(cid:19)
1
2 (cid:18) 103 cm−3
nn,0 (cid:19)
1
2 (cid:16)1 + Pext(cid:17)
1
2 pc,
[t] = 3.98 × 105(cid:18) 103 cm−3
nn,0 (cid:19)
1
2 (cid:16)1 + Pext(cid:17)
1
2 yr.
(26)
(27)
-- 14 --
The unit of mass is then
[M ] = 9.76 × 10−2(cid:18) T
10 K(cid:19)
3
2 (cid:18) 103 cm−3
nn,0 (cid:19)
1
2 (cid:16)1 + Pext(cid:17)
1
2 M⊙ .
It follows from equations (11), (12), and (25) that
τni,0 =
0.241
(1 + Pext)
2 (cid:18)3 × 10−3cm−3
K
1
(cid:19)(cid:18) 105 cm−3
nn,0 (cid:19)k−
1
2
.
From equations (24) and (25), a model's reference magnetic field is given by
Bref =
5.89 × 10−6
1
µ0(1 + Pext)
2 (cid:16)
nn,0
103 cm−3(cid:17)
1
2 (cid:18) T
10 K(cid:19)
1
2
G.
(28)
(29)
(30)
2.4. Numerical Method of Solution
The system of dimensionless partial differential equations presented in the preceding section is
solved by the method of lines (e.g., Schiesser 1991). This was used in the axisymmetric mod-
els of Morton, Mouschovias, & Ciolek (1994), Ciolek & Mouschovias (1994, 1995), and Basu
& Mouschovias (1994, 1995a, b), and the nonaxisymmetric models presented in BC04.
In this
method, spatial derivatives within the PDEs are approximated by finite differences. The square
computational domain of size L × L is discretized by dividing the region into N 2 uniform cells of
size L/N × L/N . L is typically chosen to be a factor ∼ 4 greater than the wavelength of max-
imum gravitational instability λT,m (see § 3.1 below). Three-point centered differences are used
to approximate gradients. Advection of mass and magnetic flux is performed with the monotonic
upwind scheme of van Leer (1979). The spatial discretization converts the system of PDEs to a
system of coupled ordinary differential equations of the form dy/dt = F (y, t), where dy/dt, y,
and F are all vectors of dimension V N 2 (V is the number of dependent variables). An implicit
Adams-Bashforth-Moulton method is then used to time-integrate the resulting system of ODEs.
At each time step, numerical solution of the Fourier transform and also the inverse transform of
quantities in the equations for the gravitational and magnetic potentials ([23l] and [23o]) is carried
out with standard two-dimensional Fast Fourier Transform (FFT) techniques (e.g., Press et al.
1996).
3. Stability of Cloud Models
We now consider the response of model clouds to small-amplitude disturbances or pertur-
bations. Such an analysis will illuminate the basic physics of gravitational instability in weakly
-- 15 --
ionized, sheet-like magnetic clouds, and will also provide the basis for understanding the length
and time scales over which instabilities will develop in fully nonlinear calculations.
The linear stability of isothermal self-gravitating equilibrium layers with frozen-in magnetic
fields was examined early on by Nakano & Nakamura (1978). They determined the critical mass-
to-flux ratio for such objects, which is discussed further below. The gravitational stability of
weakly-ionized and thin (but with finite thickness) axisymmetric magnetic disks was investigated
by Morton (1991) and Morton & Mouschovias (1991, unpublished). Their analysis is very similar to
the one that we present in this paper, and we reprise many of their results in the following section.
A later, independent linear study of partially-ionized magnetic disks/sheets was also presented
by Zweibel (1998). Similar to that which was done in Indebetouw & Zweibel (2000), Zweibel
(1998) studied clouds that were infinitesimally thin (Z = 0), and concentrated on those that are
magnetically subcritical.
3.1. Linearization and Analysis
As is frequently done, the unperturbed zero-order (background) state of a model is assumed
to be uniform and static. The dimensionless equations presented in § 2.3 above are linearized to
first-order by writing, for any physical quantity,
f (x, y, t) = f0 + δf
= f0 + δfaei(kxx+kyy−ωt),
(31)
where f0 refers to the unperturbed state, δf is the perturbation and δfa is its amplitude, kx and ky
are again the x and y wave numbers, and ω is the complex angular frequency. The perturbations
are small, with δfa ≪ f0. With regard to velocities, which have v0 = 0 because of the static
assumption, it is understood that δv ≪ characteristic signal speeds of the multifluid system (i.e.,
sound speed, Alfv´en speed, etc.).
For the assumed type of perturbation (31), we can set ∂/∂t → −iω, ∂/∂x → ikx, and ∂/∂y →
iky. The dimensionless equations for a model cloud then become, collecting and retaining terms to
first-order,
ω δσn = kxδvn,x + kyδvn,y ,
ω δvn,x =
ω δvn,y =
kx
ky
kz (cid:16) C 2
kz (cid:16) C 2
eff,0kz − 1(cid:17) δσn + Bref
eff,0kz − 1(cid:17) δσn + Bref
kx
kz
ky
kz
(1 + kzZ0) δBz,eq ,
(1 + kzZ0) δBz,eq ,
ω δBz,eq = Bref kxδvn,x + Bref kyδvn,y − iτni,0 B2
ref kz (1 + kzZ0) δBz,eq .
From equation (23f), the reference state effective isothermal speed of sound is
Ceff,0 = (cid:16)1 + 3 Pext(cid:17)
1 + Pext
1
2
.
(32a)
(32b)
(32c)
(32d)
(33)
-- 16 --
In deriving the linearized system (32a)-(32d), we have used equations (23l)-(23n) to relate gravi-
tational field perturbations to column density perturbations, (23o)-(23q) to relate planar magnetic
field perturbations to those of the equatorial vertical magnetic field, and (23h) and (23i) to sub-
stitute the perturbed ion velocity components with those of the neutrals. The wavenumber kz has
the same meaning as previously defined. A mode is unstable if the imaginary part of the complex
frequency ωI > 0. The growth timescale of the instability is τg = 1/ωI.
The fundamental physics of the linear system is readily discerned from the various terms in
equations (32a)-(32d): thermal-pressure and self-gravitational forces are proportional to perturba-
tions in the column density in equations (32b) and (32c), and magnetic tension and pressure forces
are proportional to perturbations in the equatorial magnetic field. The drift or diffusion of magnetic
field and plasma with respect to the neutrals is represented by the term in the magnetic induction
equation (32d) that contains τni,0; comparison with the linearized mass continuity equation (32a)
shows that, in the limit τni,0 = 0, collisional coupling of the neutrals and ions (and therefore, the
magnetic field) is instantaneous and perfect, and they all move together as a single fluid. In the
opposite extreme, τni,0 ≫ 1, neutral-ion collisions are infrequent, and the neutral and ion-magnetic
field fluids are increasingly decoupled and move independently of one another.
The solution of the full dispersion relation for the gravitationally unstable mode of the lin-
earized equations (32a) - (32d) are presented in Figure 1 for (a) τni,0 = 0, (b) τni,0 = 0.04, (c)
τni,0 = 0.1, (d) τni,0 = 0.2, (e) τni,0 = 1, and (f) τni,0 = 10. The external pressure parameter for
these models is Pext = 0.1, which sets the dimensionless effective sound speed Ceff,0 = 1.04. (Note
that Ceff,0 = 1 at Pext = 0 and 1. Ceff,0 is maximal at Pext = 1/3, and is equal to 1.061.) Displayed
in each panel of Figure 1 is the growth time τg as a function of the wavelength λ (= 2π/kz), for
various values of µ0 = 1/ Bref . The separately labeled curves show the result for µ0 = 0.5, 0.8, 1,
1.1, 2, and 10, respectively. We are able to use λ as the independent variable because the char-
1
acteristic polynomial for our eigensystem is found to be only a function of kz = (k2
2 . This
means that to this order of approximation, all perturbations are independent of the planar angle
of propagation θ [= tan−1(ky/kx)].
x + k2
y)
Understanding the data presented in Figure 1a - f is aided by examining the results for the
instability growth timescale in the following two limits.
Limit 1: Flux-freezing, τni,0 = 0.
In this limit, the neutrals, ions, and magnetic field respond as a single combined fluid, with
the magnetic flux frozen in the neutrals, and the resulting dispersion relation is found to be
The gravitationally unstable mode corresponds to one of the roots of ω2 < 0, and occurs for
refZ0(cid:17) k2
z −(cid:16) B2
ref − 1(cid:17) kz = 0.
eff,0 + B2
ω2 −(cid:16) C 2
(34)
Bref < 1, or, equivalently, µ0 > 1. The growth timescale for this mode can be written as
-- 17 --
for λ ≥ λMS, where
τg =
λ
1
2
h2π(1 − B2
λMS ≡ 2π C 2
ref )(λ − λMS)i
!
eff,0 + B2
1 − B2
ref
ref Z0
,
(35)
(36)
is the critical or threshold wavelength for instability. The minimum growth timescale (= maximum
growth rate) occurs at λMS,m ≡ 2λMS. The flux-freezing growth timescale (35) for our various
model clouds is plotted as a function of λ in Figure 1a - f , shown as open circles.
We have given to the critical wavelength λMS the subscript 'MS' because it is the maximum
lengthscale that can be supported by both magnetic and thermal pressure effects, and therefore is
related to magnetosound modes. This can be readily seen by noting that the dimensionless Alfv´en
speed in our model clouds is given by
VA,0 =
= Bref Z
1
2
0 ,
(37)
Bref
p2ρn,0
(see eq.
[23k]) and the fact that the dimensionless column density is unity for an unperturbed
cloud. It follows then that the isothermal magnetosound speed (since we consider only isothermal
perturbations in our models) in the adopted set of units is
eff,0 + V 2
VMS,0 = (cid:16) C 2
= (cid:16) C 2
= Ceff,0"1 +
eff,0 + B2
1
2
1
2
A,0(cid:17)
ref Z0(cid:17)
(1 + 3 Pext)#
(1 + Pext)
2
µ2
0
1
2
.
In the brackets of the last equality we have used eq. (33) and the relation
Z0 =
2
(1 + Pext)
,
(38)
(39)
which follows from the linearization of equations (23j) and (23k), to eliminate C 2
eff,0 and Z0 in terms
of Pext. Examination of equation (36) reveals the presence of the magnetosound speed (38) in this
expression, thus identifying the combined action of thermal and magnetic pressure in the support
of a cloud against self-gravity. It also shows the importance of magnetic pressure in setting the
instability timescale and the wavelength of maximum growth rate. For clouds that are close to
being critical (µ0 ∼ 1), neglecting this effect (e.g., Zweibel 1998; Indebetouw & Zweibel 2000) can
significantly underestimate τg and λMS,m.
-- 18 --
When there is negligible magnetic support, that is, when Bref → 0 (µ0 → ∞) it follows that
VMS,0 → Ceff,0, and λMS → λT, where
λT ≡ 2π C 2
eff ,0 = π 1 + 3 Pext
1 + Pext ! Z0
(40)
is the critical thermal lengthscale. For this situation, the growth timescale for the unstable mode
is still given by equation (35), with Bref = 0, and λT replacing λMS. The maximum growth rate
for the unstable mode in this circumstance then occurs at λT,m ≡ 2λT.
Limit 2: Stationary magnetic field lines, ωδBz,eq = 0.
For this situation, which would be relevant to models with effective ambipolar diffusion, and
therefore, relatively weak coupling of the neutrals to the ions and magnetic field, the resulting
dispersion relation is
ω2 +
i
τni,0
eff,0k2
ω −(cid:16) C 2
z − kz(cid:17) = 0.
(41)
From this relation, one finds that an unstable mode exists for λ > λT, and has a growth timescale
τg =
2τni,0λ
hλ2 + 8πτ 2
ni,0 (λ − λT)i
.
1
2 − λ
(42)
The minimum growth time for this mode is at the wavelength λT,m, defined above. The growth
time (42) is also displayed as a function of wavelength in Figure 1a - f (crosses).
As τni,0 → ∞, τg → λ/p2π(λ − λT), which is identical to the result from equation (35) when
Bref = 0. This is because in both of these circumstances the magnetic field does not affect the
neutrals: τni,0 → ∞ corresponds to when there is no collisional coupling between the neutrals
and ions (and, hence, the magnetic field). The ions are completely "invisible" to the neutrals
in this situation, and there is no transmission of magnetic force to them via neutral-ion collisions.
Similarly, when Bref = 0, there is no magnetic field, and therefore magnetic forces do not contribute
to the support or dynamics of a model cloud in that case.
With the results of the two limiting cases 1 and 2 described above in hand, the underlying
physics of the gravitationally unstable modes presented in Figure 1a - f is made more transparent.
For instance, the models with the magnetic field frozen into the neutrals (τni,0 = 0) displayed in
Figure 1a are seen to be in exact agreement with the growth timescale predicted by equation (35).
Particularly noteworthy is the dependence on the parameter µ0 (= 1/ Bref ) for these models: there
is no unstable, gravitationally collapsing mode for µ0 < 1. Moreover, models with µ0 ∼ 1 have
minimum growth timescales at wavelengths λMS,m that are much larger than those for models with
greater values of µ0. As an example, we note that λMS,m for the model with µ0 = 1.1 in Figure 1a is
an order of magnitude greater than that in the model with µ0 = 2. The actual value of the growth
timescale for the µ0 = 1.1 model is seen to be about an order of magnitude greater than that for
-- 19 --
the µ0 = 2 model as well. This is a consequence of there still being near-equality of gravitational
and magnetic forces when µ0 is very close to unity.
When there is imperfect neutral-ion coupling (τni,0 > 0), gravitational instability is possible
even for model clouds with µ0 < 1. This can be seen in panels b - f of Figure 1. The timescale for
the instability is typically greater than that for the supercritical flux-frozen models (Fig. 1a), but
it is finite. For these models, a Jeans-like growth of density perturbations occurs with the collapse
moderated by the retarding collisional forces exerted on the neutrals as they diffuse through the
plasma and field. A result of this type was first noted by Langer (1978). Models that are more
subcritical are better approximated by the stationary field limit described above. This can be seen
by the excellent agreement of the limiting growth timescale given by equation (42) -- and also
the wavelength of maximum growth (λT,m) -- with the dispersion curve for the µ0 = 0.5 model
displayed in Figure 1b - f . As µ0 approaches unity, there is a transition of the growth timescale
behavior from the stationary field/ambipolar diffusion limit (eq. [42]) and toward the flux-freezing
limit (eq.
[35]). For µ0 close to unity, the instability proceeds through a hybrid mode in which
both ion-neutral drift and gravitational contraction with field-line dragging are active. As discussed
above, the two limiting approximations approach one another for models with τni,0 ≫ 1 and µ0 ≫ 1,
since either limit describes clouds with effectively no magnetic support. This accounts for the near
equality of all the model curves seen in Figure 1f .
Figure 2 presents the wavelength λg,m, which is defined as the wavelength λ that has the
minimum growth time, as a function of µ0 for models with the same values of τni,0 and Pext as
in Figure 1a - f . For comparison, the dashed line in Figure 2 is the constant value of λT,m for
those particular models. Consistent with our discussion above, we note that there is singular and
limiting behavior of this lengthscale for the model cloud with τni,0 = 0 and µ0 = 1. The value of
λg,m is also seen to be especially sensitive to the value of µ0 for near-critical clouds with τni,0 > 0,
exhibiting a sharp, resonant-like peak in the region µ0 ∼ 1. Table 1 lists the ratio λg,m/λT,m at
the peak for each of the non-flux-frozen models in Figure 2. The peak ratio is largest for models
with τni,0 . 1, while, for models with τni,0 ≫ 1, magnetic field effects are barely transmitted to the
neutrals (because neutral-ion collisions are very rare), and this ratio instead approaches unity.
Based on these results, we expect then, that clouds that are marginally or slightly supercritical
will form cores that evolve more slowly, and have size scales (radii, and core spacing) significantly
greater than that which will occur in clouds that are more highly supercritical. In addition, nearly
critical clouds will also have size scales that are markedly larger than those in distinctly subcritical
clouds. As we shall see in a following nonlinear study, this specific dependence of the gravitational
instability on the initial mass-to-flux ratio leads to notable differences in the physical characteristics
of collapsing cores in magnetic interstellar clouds that should be easily discerned by observations.
Figure 3a - f shows the growth timescales for the same model clouds as in Figure 1, but with
the dimensionless external pressure Pext = 10. In the limit of large Pext, equation (23j) shows that
the cloud is pressure confined, with Z ∝ σn (eq. [23k]). As a result, a local peak in σn is a peak in Z.
-- 20 --
For this situation then, the top surface of the cloud looks like a dome and the external pressure force
has horizontal components that point inward to the dome's peak, acting in the same directions as
the gravitational field components gx and gy. This further enables the gravitational clumping, and
decreases the growth timescale, appearing as a reduction of Ceff,0 in our equations. Because of this,
the dimensionless sound speed Ceff,0 is decreased by a factor of 2.05 from its value in the models
with Pext = 0.1. This also reduces the growth timescale τg and associated lengthscales λMS and λT,
since they are also functions of Ceff,0, and therefore, of Pext. The net effect is to shift and reduce
the instability timescale and critical wavelengths by a corresponding factor from those seen in the
models with lower external pressure. This means that clouds in regions with much greater external
pressure (perhaps due to being embedded in a massive cloud complex or fragment, or adjoining
hotter environments such as an HII region and/or shocked gas) will have characteristic sizes that
are smaller when compared to clouds surrounded by lesser external pressures. However, the general
behavior and physics of the models as a function of τni,0 and µ0, including the applicability of
the limiting analytical approximations discussed above, is similar to that seen in the previously
described models with smaller Pext.
The wavelength with the minimum growth time λg,m is shown as a function of µ0 in Figure 4
for models with the same values of τni,0 and Pext as in Figure 3. It is again the case that, for clouds
that are almost critical and with τni,0 . 1, the wavelength that has the most rapid gravitational
response is significantly greater than in models that are farther away from µ0 = 1. The peak ratio
values of λg,m to λT,m for the models with τni,0 > 0 are also listed in Table 1. As before, this ratio
is much greater for the models with τni,0 . 1.
3.2. Comparison to numerical simulations
We now compare the results of the dispersion analysis of the preceding section to simulations of
the early evolution of model clouds governed by the full system of non-ideal magnetohydrodynamic
equations (23a) - (23q). This system of equations is solved by a numerical code using the techniques
described in § 2.4. The code was previously used to generate results for two model clouds whose
nonlinear evolution was presented by BC04. During the early evolution of a model cloud, physical
variables do not change very much from their initial state values. Hence, during this period the
physical evolution should be close to that determined by the linear analysis above. Comparing the
results of the early-time evolution of the full simulations to those of the dispersion analysis allows
us to reinforce our understanding of the underlying physics governing the early evolution of a cloud
-- the precursor to the later nonlinear phase of evolution -- as well as provide a useful benchmark
and theoretical standard to establish the overall accuracy of our numerical code.
Because we are concerned with the linear stages of gravitational instability, we will confine our
focus of the numerical simulations in this study to the evolution of the column density in model
clouds. The detailed time-dependent behavior of other quantities, such as the velocity fields of
the ions and neutrals, and the evolution and redistribution of mass in magnetic flux tubes will be
described in a following paper.
-- 21 --
3.2.1. Monochromatic perturbations
We follow the evolution of clouds initially given perturbations with a single wavelength λ. The
evolution of a model cloud is initiated by imposing at time t = 0 an initial (normalized) column
density profile of the form
σn(x, y, 0) = 1 + δσn,a cos(2πx/λ),
(43)
that is, a uniform background state with a perturbation with amplitude δσn,a.
In using this
particular perturbation, we make use of the fact that, as mentioned in § 3.1 above, the dispersion
analysis indicates that the effects of linear disturbances are independent of the angle of propagation
θ. Since it really doesn't matter for our purposes which direction of propagation we choose, we
take θ = 0 (parallel to the x-axis), which means that ky = 0, and therefore kx = kz = 2π/λ. The
initial velocity and magnetic field are also perturbed in a way that is consistent with the system of
equations (32a) - (32d) for the column density perturbation specified by equation (43). Solving for
the initial perturbations δvn,x, δBz,eq, and δvn,y in terms of the given δσn, kx, and τg = i/ω (which
is a function of λ) from this set of equations yields
δvn,x(x, y, 0) = −
δBz,eq(x, y, 0) =
λ
2πτg
δσn,a sin(2πx/λ),
λ Bref
λ + 2πτg τni,0 B2
ref (1 + 2πZ0/λ)
δσn,a cos(2πx/λ) ,
(44)
(45)
and δvn,y(x, y, 0) = 0. By specifying the relation between the perturbed physical variables in this
way, we are basically selecting the eigenvector of the perturbation at the wavelength λ. Hence, the
monochromatic perturbation excites a single eigenmode of a model cloud at t = 0.
Not surprisingly, when we initiate the time evolution in this fashion, the subsequent evolution
of a model cloud is simply the continued growth of the excited lone eigenmode. The left panel
of Figure 5 displays the growth of the column density maxima for four different cloud models as
a function of time. Each has τni,0 = 0.2 and Pext = 0.1. For all of these models, δσn,a = 0.02
and λ = 4π. The calculations were performed on an equally spaced mesh of 32 × 32 cells on
a square computational region of size L = 4π in each direction. Shown in that Figure (solid
curves) is the evolution of the maximum reduced column density δσn(t) = σn(t) − 1 for models
with µ0 = 0.5, 1, 2, and 10. Also shown (dashed curves) is the growth of the column density
perturbations that would occur as predicted by the linear relation for an unstable mode at the
same location, δσn(t) = δσn,a exp(t/τg). The values of τg used to plot the curves are those derived
from the earlier dispersion analysis for λ = 4π, τni,0 = 0.2 and Pext = 0.1. From the data in Figure
1d we find that τg = (20.3, 16.0, 4.34, 2.12) for µ0 = (0.5, 1.0, 2, 10). Comparing the numerical
simulation results to the theoretical values predicted by the linear analysis indicates that they are
-- 22 --
in excellent agreement and overlap during much of the early evolution of each model.
In fact,
significant deviation (& 1%) between the simulation results and the linear predictions does not
begin to occur until δσn has grown to ∼ 0.2, which is well beyond what is generally considered
the regime of linear growth. At later times nonlinear effects are evident, and the column density
grows more rapidly and significantly exceeds that predicted by the linear analysis by the end of the
period shown in Figure 5, when δσn = 1.
The growth of the column density perturbations in the numerical simulations at early times
is entirely consistent with our discussion of the physical processes acting in the dispersion analysis
in § 3.1. For instance, the linear analysis predicts that the onset of gravitational instability in the
subcritical µ0 = 0.5 model is due to the action of ambipolar diffusion (see Fig. 1d). Moreover,
using the stationary field limit approximation (eq. [42]) to calculate τg for this model (λT = 6.75
for these parameters) gives τg = 21.8, which differs from the full dispersion analysis result quoted
above by only 7%. The critical model µ0 = 1 and supercritical model µ0 = 2 are also heavily
influenced by ambipolar diffusion, and they would not be able to collapse in its absence, because
the excited wavelength for both of these models is in the region λT < λ < λMS. Thus, these
two models lie in the transition region between diffusion-regulated and flux-frozen collapse for this
particular wavelength, as noted in our discussion of these models in regard to Figure 1d. Finally,
the instability of the highly supercritical model µ0 = 10 occurs essentially with freezing of magnetic
field lines in the neutral matter. For this model, λMS = 6.93 which is < 4π. Using the flux-frozen
approximate expression for the growth time given by equation (35) yields τg = 2.12 for this model,
which is exactly the same as the full dispersion relation solution, and is also in very good agreement
with the model simulation's early-time linear growth.
The right panel of Figure 5 shows the growth of the column density maxima in another four
model clouds, with the same wavelength and parameters except Pext = 10. The development of
these higher- Pext models is again seen to be in excellent agreement with the dispersion analysis for
the linear stage of evolution. They are also in accordance with the values of τg predicted by the
analytic relations (42) and (35) for the subcritical and supercritical models, respectively.
We conclude then, that our numerical code accurately reproduces the physics of gravitational
instability in planar magnetic model clouds.
3.2.2. White noise perturbations: the wavelength of minimum growth time λg,m and its
dependence on µ0
In this subsection we consider cloud models that are given a spectrum of random, small-
amplitude perturbations in the physical variables at t = 0. The spectrum we use is white noise, i.e,
flat, so that there is no preferred wavelength selected by the perturbations. However, we introduce
damping so that wavelengths equal to twice the mesh spacing and smaller are negligible. The
root-mean-square amplitude of the fluctuations in the initial state of a model cloud is 3% of the
uniform background. The size of the computational region in each direction is L = 16π, and the
-- 23 --
number of cells in each direction is N = 128.
The white noise spectrum will excite an ensemble of eigenmodes with a wide range of wave-
lengths. The dominant evolutionary modes that will emerge from this ensemble of fluctuations and
govern the subsequent evolution will be those grouped about the one with the minimum growth
time. A cloud will then develop features with lengthscales comparable to the wavelength with the
minimum growth time, λg,m. That this intuitive notion is valid can be seen in Figure 6, which
shows contour maps of the column density σn(x, y) in two model evolution simulations at the time
for which the value of the maximum column density has grown to σn,max = 2 (twice the initial
background value) in each model. Both models again have τni,0 = 0.2 and Pext = 0.1. The model
displayed in the left panel is initially subcritical with µ0 = 0.5, and the one in the right panel started
as a critical cloud with µ0 = 1. The density contours cover the range 0.8 to 1.6, in increments of 0.2.
By the time that the contour snapshots are taken, nonlinear effects have set in. This is evidenced
by column densities that are relatively large compared to that of the initial background state, and
the beginning of nonaxisymmetric gravitational fragmentation (which will form protostellar cores)
due to the interaction of various eigenmodes. Despite this, both cloud models retain a significant
imprint from the predecessor linear stage of evolution: there is a nearly uniform separation between
core fragments -- defining here a core as the enclosed higher column density regions with σn > 1.2
-- and the fragmentation scale for these nascent cores is close to λg,m in each model. For the model
with µ0 = 0.5, Figure 2 indicates that λg,m = 14, while for the cloud with µ0 = 1, λg,m = 24. Exam-
ination of Figure 6 reveals that these scales correspond to the mean distance between the different
σn = 1.2 contours for these models. Additionally, we note that the behavior of the fragmentation
scale in these models as a function of µ0 is also consistent with the linear analysis. Notably, the
average spacing between cores in the critical model evolution simulation is larger than that in the
subcritical model, in line with the predicted dependence of λg,m on µ0 (see Fig. 2). This results in
there being fewer cores within the same size region. In fact, according to Table 1 (see, also, Fig.
1d), the maximal λg,m(= 52.8) for these values of τni,0 and Pext occurs at µ0 = 1.13.
Further confirmation that λg,m sets the characteristic lengthscales in clouds can be seen in
Figure 7, which presents the column density plots of two more model clouds, with the same values
of τni,0 and Pext as in Figure 6. The model in the left panel of Figure 7 is supercritical and has
µ0 = 2. The linear analysis predicts λg,m = 23 for this model, which is indeed close to the the
distance between core-bounding contours in that Figure. There are also relatively fewer cores that
have formed, similar to that seen in the critical model. The similarity in the structural (lengthscale)
properties of the µ0 = 1 and µ0 = 2 models is not surprising, as they have nearly equal values for
λg,m. However, the dynamical properties of the µ0 = 1 and µ0 = 2 models are quite different in
the nonlinear regime, as shown by BC04. Specifically, the maximum velocities in the supercritical
models become supersonic on scales that are well within the resolving power of modern observations.
We defer the detailed discussion of the dynamical and related nonlinear evolution of cloud models
to a study that will follow this paper, and to that already presented in BC04. Finally, the right
panel of Figure 7 shows the column density contours for a highly supercritical model with µ0 = 10.
-- 24 --
For this model, λg,m = 13. This is also in agreement with the mean distance between cores in the
contour plot. The lengthscales and number of cores in the highly supercritical model are akin to
that seen in the subcritical µ0 = 0.5 model. This is no coincidence, since these two models have
similar values of λg,m. Although a somewhat counterintuitive result, this is because there is also
a Jeans-like growth of density perturbations in the subcritical models, but it occurs on the much
longer ambipolar diffusion timescale as neutrals drift past near-stationary plasma and magnetic
field.
There are also some differences in the models in Figures 6 and 7 that become more pronounced
with increasing initial mass-to-flux ratio: the cores become more nonaxisymmetric with greater
values of µ0, showing enhanced elongation along a single axis. This second-order effect -- an
amplification of nonaxisymmetry due to the nonlinear interaction of certain eigenmodes for the
more supercritical models -- will be a topic of investigation in a following paper. And, as mentioned
above, the dynamical properties of the model clouds, such as their velocity fields, are also markedly
different, characteristically having greater infall velocities over larger scales with increasing µ0.
This too will be studied further in a future paper. Despite these differences, we conclude from our
representative evolution simulations that the gravitational fragmentation scale of a model cloud
with an initial spectrum of perturbations is effectively the value of λg,m for that particular model.
The fragmentation scale has a dependence on µ0 in agreement with the result of the linear analysis,
especially with the prediction that near the critical value there is a dramatic increase in λg,m.
Because of this resonant-like behavior, clouds that are close to critical will have significantly larger
size scales than in models that are more highly supercritical or more highly subcritical.
4. Summary
We have presented the formulation of physical models that we will use to study the nonax-
isymmetric formation and self-gravitational collapse of protostellar cores in magnetic interstellar
molecular clouds. Model clouds are partially ionized isothermal thin planar sheets with finite
half-thickness Z(x, y).
The system of equations that are used to govern the time evolution of model clouds contain four
fundamental parameters, all defined in § 2.3. The first is the dimensionless background magnetic
field strength of a cloud, Bref ; the inverse of this parameter is the initial mass-to-magnetic flux
ratio in units of the critical value for gravitational collapse, µ0. Clouds with Bref > 1 (µ0 < 1)
are subcritical and are initially magnetically supported, and cannot collapse in the absence of
ambipolar diffusion. Models with Bref < 1 (µ0 > 1) are supercritical from the outset and unable to
support themselves against their own self-gravity. The next parameter is the normalized neutral-ion
collisional momentum-exchange time τni,0. It is a measure of the efficiency of ambipolar diffusion in
model clouds: τni,0 → 0 corresponds to very effective momentum transfer and collisional coupling
of neutrals and ions, resulting in freezing of magnetic flux in the neutral matter. In the opposite
limit of τni,0 ≫ 1, collisions are so infrequent that the neutrals are essentially decoupled from the
-- 25 --
plasma and magnetic field, and magnetic forces contribute negligibly to the support and evolution
of a model cloud. Pext is the ratio of the external pressure to the vertical self-gravitational stress
in the initial uniform reference state of a model cloud. The final parameter is the exponent k in
the power-law expression used to calculate the ion density as a function of the neutral density.
We also investigated the linear stability of model clouds, and how the growth times τg and
critical wavelengths of the gravitationally unstable modes depend on µ0, τni,0, and Pext. Analytic
expressions for τg and the critical wavelengths were derived for the limits of weak and strong
ambipolar diffusion. These expressions (eqs.
[35] and [42]) agreed well with the full dispersion
results in these limits.
Models with frozen-in magnetic flux (τni,0 = 0) are gravitationally unstable when they are
supercritical and the wavelength of a perturbation exceeds the magnetosound critical wavelength
λMS. For finite thickness clouds, magnetic pressure contributes substantially to setting the value
of λMS and τg when µ0 ∼ 1. Ambipolar diffusion (τni,0 > 0) allows clouds to be unstable even
when subcritical, so long as the perturbation's wavelength is greater than the thermal critical
wavelength λT (λT ≤ λMS, see eqs. [36] and [40]). The instability modes of clouds with Pext > 1
behave qualitatively the same way as for those with Pext ≤ 1, except that quantitatively their
critical wavelengths and growth times are reduced, due to the effective retarding sound speed being
decreased at higher external pressures. The dispersion analysis also revealed how the wavelength
with the minimum growth time, λg,m, behaves as a function of µ0. For models with τni,0 > 0, λg,m
has a resonance at a value of µ0 that is usually just slightly greater than the critical value µ0 = 1.
Because of this resonance, a model cloud with µ0 ∼ 1 will have a λg,m that is significantly greater
when compared to that of models that are much more sub- or supercritical.
In addition to providing basic insight to the physics of gravitational instability in partially
ionized media, the linear analysis was also used as a theoretical standard to test the accuracy of
our numerical method of solution of the full set of nonlinear equations (eqs.
[23a] - [23q]) that
govern the time evolution of a model cloud. The early-time evolution of several cloud simulations
given an initial monochromatic perturbation of wavelength λ was compared to the predicted value
of the growth timescale from the dispersion analysis. The temporal growth of the column density
maxima in each model simulation was in excellent agreement with the theoretically predicted values
of τg, thus establishing the accuracy of our non-ideal magnetohydrodynamic computational code.
Finally, we presented a suite of model simulations that had their evolution initiated by fluctu-
ations with a spectrum of wavelengths, following their evolution a little beyond the linear growth
phase. The characteristic fragmentation scale that developed in these models tends to correspond
to the wavelength with the minimum growth time λg,m of the initial state. The resonance behavior
of λg,m for clouds with the parameter µ0 near the critical value -- as predicted by the linear analysis
-- was also in evidence in the numerical simulations. Those having µ0 ∼ 1 had a much greater
mean spacing (≈ λg,m) between cores than in models with µ0 ≪ 1 and µ0 ≫ 1. Because of this, the
total number of cores that developed in the models near the resonance was less than in those that
-- 26 --
were farther away. This sensitive dependence of λg,m about µ0 ∼ 1 has important implications for
core and star formation, since observations currently indicate that mass-to-flux ratios in molecular
clouds generally lie in the range 0.5 . µ0 . 2 (Crutcher 2004).
In forthcoming studies we will explore further the formation and nonaxisymmetric collapse of
protostellar cores as a function of the fundamental model parameters, building on the formulation
and analysis presented in this paper. The properties of dynamically infalling cores such as spatial
density and velocity maps, core shapes, magnetic field strengths, and other quantities of interest,
will be investigated in detail. In doing so, we will be providing a physically consistent model with
testable, quantitative predictions that may be used to interpret and perhaps guide observations of
star formation in magnetic interstellar molecular clouds.
GC was supported by the New York Center for the Origins of Life (NSCORT) and the De-
partment of Physics, Applied Physics, and Astronomy at Rensselaer Polytechnic Institute, under
NASA Grant NAG5-7589. SB was supported by a grant from the Natural Sciences and Engi-
neering Research Council of Canada. Helpful comments from an anonymous referee are gratefully
acknowledged.
-- 27 --
Table 1
Peak ratio of wavelength with minimum growth time (λg,m) to wavelength of
maximum gravitational instability (λT,m)
τni,0
µ0
(cid:16) λg,m
λT,m(cid:17)peak
Pext = 0.1:
Pext = 10:
0.04
0.1
0.2
1
10
0.04
0.1
0.2
1
10
1.02
1.06
1.13
1.64
3.50
1.06
1.11
1.23
1.87
3.31
21.1
8.37
4.20
1.42
1.04
6.04
2.89
1.90
1.17
1.01
-- 28 --
REFERENCES
Allen, A., & Shu, F. H. 2000, ApJ, 536, 368
Andr´e, P., Motte, F., & Belloche, A. 2001, in From Darkness to Light: Origin and Evolution of
Young Stellar Clusters, ed. T. Montmerle & P. Andr´e (San Francisco: ASP), 209
Bacmann, A., Andr´e, P., Puget, J.-L., Abergel, A., Bontemps, S., & Ward-Thompson, D. 2000,
A&A, 361, 555
Basu, S. 1997, ApJ, 485, 240
. 2000, ApJ, 540, L103
Basu, S., & Ciolek, G. E. 2004, ApJ, 607, L39 (BC04)
Basu, S., & Mouschovias, T. Ch. 1994, ApJ, 432, 720 (BM94)
. 1995a, ApJ, 452, 386
. 1995b, ApJ, 453, 271
Byron, F. W., Jr., & Fuller, R. W. 1992, Mathematical Methods of Classical and Quantum Physics
(Mineola: Dover)
Caselli, P., Walmsley, C. M., Zucconi, A., Tafalla, M., Dore, L., & Myers, P. C. 2002, ApJ, 565,
331
Chandrasekhar, S., & Fermi, E. 1953, ApJ, 118, 113
Ciolek, G. E., & Basu, S. 2000, ApJ, 529, 925 (CB00)
Ciolek, G. E., & Mouschovias, T. Ch. 1993, ApJ, 418, 774 (CM93)
. 1994, ApJ, 425, 142
. 1995, ApJ, 454, 194
. 1996, ApJ, 468, 749
. 1998, ApJ, 504, 280
Ciolek, G. E., & Konigl, A. 1998, ApJ, 504, 257
Contopoulos, I., Ciolek, G. E., & Konigl, A. 1998, ApJ, 504, 247
Crapsi, A., Caselli, P., Walmsley, C. M., Tafalla, M., Lee, C. W., Bourke, T. L., & Myers, P. C.
2004, A&A, 420, 957
-- 29 --
Crutcher, R. M. 2004, in The Magnetized Interstellar Medium, ed. B. Uyaniker, W. Reich, & R.
Wielebinski (Katlenburg-Lindau: Copernicus GmbH), 123
Crutcher, R. M., Nutter, D. J., Ward-Thompson, D. W., & Kirk, J. M. 2004, ApJ, 600, 279
Crutcher, R. M., & Troland, T. H. 2000, ApJ, 537, L139
Curran, R. L., Chrysostomou, A., Collett, J. L., Jenness, T., & Aitken, D. K. 2004, A&A, 421, 195
Fiedler, R. A., & Mouschovias, T. Ch. 1993, ApJ, 415, 680
Fujiyoshi, T., Smith, C. H., Wright, C. M., Moore, T. J. T., Aitken, D. K., & Roche, P. F. 2001,
MNRAS, 327, 233
Indebetouw, R. M., & Zweibel, E. G. 2000, ApJ, 532, 361
Jones, C. E., & Basu, S. 2002, ApJ, 569, 280
Jones, C. E., Basu, S., & Dubinski, J. 2001, ApJ, 551, 387
Kerton, C. R., Brunt, C. M., Jones, C. E., & Basu, S. 2003, A&A, 411, 149
Kiguchi, M., Narita, S., Miyama, S., & Hayashi, C. 1987, ApJ, 317, 830
Kudoh, T., & Basu, S. 2003, ApJ, 595, 842
. 2006, ApJ, 642, 270
Lai, S.-P., Crutcher, R. M., Girart, J. M., & Rao, R. 2001, ApJ, 561, 864
. 2002, ApJ, 566, 925
Langer, W. D. 1978, ApJ, 225, 95
Li, Z.-Y., & Nakamura, F. 2004, ApJ, 609, L83
Li, Z.-Y., & Shu, F. H. 1997a, ApJ, 475, 237
. 1997b, 475, 251
McDaniel, E. W., & Mason, E. A. 1973, The Mobility and Diffusion of Ions and Gases (New York:
Wiley)
Morton, S. A. 1991, Ph.D. thesis, University of Illinois, Urbana-Champaign
Morton, S. A., Mouschovias, T. Ch., & Ciolek, G. E. 1994, ApJ, 421, 561
Motte, F., Andr´e, P., & Neri, R. 1998, A&A, 336, 150
Mouschovias, T. Ch. 1976, ApJ, 207, 141
-- 30 --
. 1977, ApJ, 211, 147
. 1978, in Protostars and Planets, ed. T. Gehrels (Tucson: U. Arizona Press), 209
. 1979, ApJ, 228, 475
Mouschovias, T. Ch., & Ciolek, G. E. 1999, in The Origin of Stars and Planetary Systems, ed. C.
J. Lada & N. D. Kylafis (Dordrecht: Kluwer), 305
Nakamura, F., & Li, Z.-Y. 2005, ApJ, 631, 411
Nakano, T., & Nakamura, T. 1978, PASJ, 30, 671
Narita, S., Hayashi, C., & Miyama, S. 1984, Prog. Theo. Phys., 72, 1118
Pereyra, A., & Magalhaes, A. M. 2004, ApJ, 603, 584
Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1996, Numerical Recipes in
Fortran 77: The Art of Scientific Computing (Vol. 1 of Fortran Numerical Recipes), 2nd.
Ed. (New York: Cambridge)
Schiesser, W. E. 1991, The Numerical Method of Lines: Method of Integration of Partial Differential
Equations (San Diego: Academic)
Schleuning, D. A. 1998, ApJ, 493, 811
Schleuning, D. A., Vaillancourt, J. E., Hildebrand, R. H., Dowell, C. D., Novak, G., Dotson, J. L.,
& Davidson, J. A. 2000, ApJ, 535, 913
Spitzer, L., Jr. 1978, Physical Processes in the Interstellar Medium (New York: Wiley-Interscience)
Tafalla, M., Mardones, D., Myers, P. C., Caselli, P., Bachiller, R., & Benson, P. J. 1998, ApJ, 504,
900
Tassis, K., & Mouschovias, T. Ch. 2005a, ApJ, 618, 769,
. 2005b, ApJ, 618, 783
van Leer, B. 1979, J. Comput. Phys., 32, 101
Williams, J. P., Myers, P. C., Wilner, D. J., & Di Francesco, J. 1999, ApJ, 513, L61
Wyld, H. W. 1976, Mathematical Methods for Physics (Reading: Benjamin)
Zucconi, A., Walmsley, C. M., & Galli, D. 2001, A&A, 376, 650
Zweibel, E. G. 1998, ApJ, 499, 746
This preprint was prepared with the AAS LATEX macros v5.2.
-- 31 --
Fig. 1. -- Instability growth times for various model clouds as a function of wavelength, for models
with (a) τni,0 = 0, (b) τni,0 = 0.04, (c) τni,0 = 0.1, (d) τni,0 = 0.2, (e) τni,0 = 1, and (f) τni,0 = 10,
respectively.
In each panel are shown the timescale curves for models with mass-to-flux ratios
µ0 = 0.5, 0.8, 1, 1.1, 2, and 10 (labeled). Each model has Pext = 0.1. Also shown are the results
of the approximate analytical solutions in the limit of flux-freezing (open circles), and stationary
magnetic field lines with ambipolar diffusion (crosses), given by eqs. (35) and (42), respectively.
-- 32 --
Fig. 2. -- Wavelength with the minimum growth time (= maximum growth rate) as a function
of initial mass-to-flux ratio. The displayed curves are for models with τni,0 = 0, 0.04, 0.1, 0.2, 1,
and 10, respectively. All models have Pext = 0.1. The dashed curve is the value of λT,m for these
models.
-- 33 --
Fig. 3. -- Same as in Fig. 1, but with Pext = 10.
-- 34 --
Fig. 4. -- Same as in Fig. 2, but with Pext = 10.
Fig. 5. -- Time evolution of column density maxima for numerical simulations of various cloud
models given an initial monochromatic perturbation of wavelength λ = 4π. The amplitude of
the initial perturbation of the column density is δσn,a = 0.02. Each model has τni,0 = 0.2, and
is labeled by the value of its mass-to-flux ratio µ0, which is 0.5, 1, 2, and 10, respectively. Left
panel: models with Pext = 0.1. Right: models with Pext = 10. The solid lines show the peak
density in each model simulation. The dashed lines are the theoretical values given by the relation
δσn(t) = δσn,a exp(t/τg), where τg is the growth time for each model predicted from the linear
analysis.
-- 35 --
Fig. 6. -- Column density contours of two model clouds that initially had random small-amplitude
fluctuations in physical variables superposed on a uniform background state. The contours are
overlaid on a grayscale image of the logarithm of the column density. At the time shown, the peak
column density in each model is twice the value of the background state (σn,max = 2). Each model
has τni,0 = 0.2 and Pext = 0.1. The left model has µ0 = 0.5, and the right one has the critical value
µ0 = 1. The density contours range from 0.8 to 1.6, in steps of 0.2.
Fig. 7. -- Same as in Fig. 6, but for supercritical models with µ0 = 2 (left) and µ0 = 10 (right).
|
astro-ph/9712232 | 1 | 9712 | 1997-12-17T15:42:32 | The viscous disk of GRS 1915+105 | [
"astro-ph"
] | GRS 1915+105, one of the two known galactic microquasars, shows an extremely complex variability in the X-ray band, comparable to no other X-ray source in the sky. Making use of RXTE/PCA data, we have analyzed the X-ray spectral distribution throughout the variability. We find that all variations can be attributed to the rapid appearance and disappearance of the inner region of an optically-thick accretion disk. Since the time scale for each event is related to the maximum radius of the disappearing region, the difference in time structure is due to the time distribution of such radii. The observed relation between the extent of the missing inner region of the disk and the duration of an event is in remarkable agreement with the expected radius dependence of the viscous time scale of the radiation-dominated region of an accretion disk. | astro-ph | astro-ph |
The viscous disk of GRS 1915+105
T. Bellonia ∗, M. M´endeza,b, A.R. Kingc, M. van der Klisa, J. van Paradijsa,d
aAstronomical Institute "A. Pannekoek", University of Amsterdam and Center for High-Energy
Astrophysics, Kruislaan 403, NL-1098 SJ, Amsterdam, The Netherlands
bFacultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata, Paseo del Bosque
S/N, 1900 La Plata, Argentina
cAstronomy Group, University of Leicester, Leicester LE1 7RH, United Kingdom
dPhysics Department, University of Alabama in Huntsville, Huntsville, AL 35899, USA
GRS 1915+105, one of the two known galactic microquasars, shows an extremely complex variability in the
X-ray band, comparable to no other X-ray source in the sky. Making use of RXTE/PCA data, we have analyzed
the X-ray spectral distribution throughout the variability. We find that all variations can be attributed to the
rapid appearance and disappearance of the inner region of an optically-thick accretion disk. Since the time scale
for each event is related to the maximum radius of the disappearing region, the difference in time structure is due
to the time distribution of such radii. The observed relation between the extent of the missing inner region of the
disk and the duration of an event is in remarkable agreement with the expected radius dependence of the viscous
time scale of the radiation-dominated region of an accretion disk.
1. INTRODUCTION
GRS 1915+105 is the first galactic object for
which radio jets moving at superluminal speed
have been observed [1]. Radio observations have
also led to the estimate of its distance (∼12.5 kpc)
and the inclinations of the jets (70◦ from the line
of sight). The source is considered to host a black
hole because of its high X-ray luminosity and its
similarities to GRO 1655 -- 40, whose dynamical
mass estimate indicate the presence of a black
hole [2]. GRS 1915+105 is a transient source,
in the sense that it has made its appearance in
the X-ray sky in 1992 [3], but most likely never
returned into quiescence since then. The Rossi
X-ray Timing Explorer (RXTE) monitors regu-
larly GRS 1915+105 since the start of the mis-
sion. During this period, the source has shown
a remarkable richness in variability, ranging from
quasi-periodic burst-like events, deep regular dips
and wild oscillations, alternated with quiescent
periods [4 -- 9]. We present here the results of the
∗This work is partly supported by NWO under grant PGS
78 -- 277
analysis of RXTE data of GRS 1915+105 and
propose an interpretation for the variability in the
spectral parameters. A more detailed discussion
can be found in [8].
2. X-RAY OBSERVATIONS
RXTE observed GRS 1915+105 in many occa-
sions. Here we present the results of the analysis
of one of these observations, the one obtained on
1997 June 18 starting at 14:36 UT and ending at
15:35 UT, as it reproduces within one day most
of the variability observed from this source. The
upper panel of Figure 1 shows 1200 s of the 2-
40 keV light curve. It consists of a sequence of
'bursts' of different duration with quiescent in-
tervals in between. All bursts start with a well-
defined sharp peak and decay faster than they
rise. The longer bursts show oscillation (or sub-
bursts) towards the end.
The complexity of the light curve seems to be
untractable: the only obvious comment that can
be made is that it consists of high flux intervals
In
somehow alternated with low flux intervals.
2
order to quantify the timing structure of the os-
cillations, we measured the length of the 'ups'
and 'downs', an unambiguous procedure given the
square-wave structure of the light curve. Exclud-
ing the fastest oscillations where this measure-
ment is not easy, we obtain durations spanning
from a few seconds to three minutes.
To study the evolution of the spectrum of
GRS 1915+105 during these oscillations between
'up' and 'down', we first produced a color-color
(C-C) diagram of the whole observation A char-
acteristic pattern emerged, pattern which is also
observable, with small differences, in all other ob-
servations we analyzed, besides a few (see below).
This strengthens the hypothesis that all the os-
cillations have the same nature. An analysis of
separate C-C diagrams for oscillations of differ-
ent length (see Figure 2) shows systematic differ-
ences. As it can also be seen from the light curve,
the 'ups' are very similar to each other, while the
'downs' are different, in particular the longer the
oscillation the deeper (light curve) and harder (C-
C) is the 'down'. Both from the light curve and
the C-C diagram it appears that the short 'low'
periods are similar to the last part of the long
ones. Since spectral analysis at the required time
resolution is not possible, in order to transform
the information in the C-C diagram into spectral
parameters we divided it into small regions. For
all the populated regions in the diagram we accu-
mulated 1 second time resolution energy spectra
in 48 energy bands. We measured a background
spectrum from a blank sky observation, which we
normalized to the highest energy channels where
the contribution of the source was negligible. We
subtracted this background spectrum from each
of the source energy spectra. We used the latest
detector response matrix available, and added a
systematic error of 2% to account for the cali-
bration uncertainties. For each of the regions in
the C-C diagram we fitted the data with a "stan-
dard" spectral model for BHCs, consisting of a
disk-blackbody (DBB) model and a power law,
both affected by interstellar absorption. To avoid
problems due to the background subtraction, and
as we were only interested in the properties of the
DBB component, we limited our fits to energies
below 30 keV. Since both the distance to this sys-
tem and the inclination of the accretion disk are
known (we assume that the jet is perpendicular
to the disk), we can derive the inner radius of
the accretion disk directly from the fits without
significant additional uncertainties.
This rather complicated procedure allowed us
to obtain time histories for the inner radius and
the temperature of the disk (bottom two panels of
Figure 1). During 'ups' the temperature is above
2 keV and the radius is stable around 20 km. Dur-
ing 'downs', the temperature drops to less than
1 keV and the radius increases.
We produced similar C-C diagrams
for
a number of other RXTE observations of
GRS 1915+105. All the observations that we an-
alyzed can be fitted in the same manner, except
that of 1996 June 16th [4] (and some later ob-
servations in October 1997). The quasi-periodic
bursts observed in many of the observations [9]
are consistent with repetitive short events like the
ones described here.
This analysis brought us to consider more se-
riously the 'up' and 'down' paradigm: to be con-
sistent with our previous work, we will call them
'outburst' and 'quiescent' states. A sequence of
quiescence and outburst we call an 'event'.
3. DISCUSSION
The main conclusion that can be derived from
Figure 1 is that the inner radius of the accretion
disk is not constant in time. If we interpret the
small and stable radius of ∼20 km observed dur-
ing outburst as the minimum stable orbit around
a black hole (see [7] for a more detailed discus-
sion), this implies that during quiescence the cen-
tral section of the disk disappears, i.e. a central
hole appears in the disk. The radius of this cen-
tral hole varies between events, ranging in this
observation between 30 and 90 km (see Figure
1). This can be interpreted within the model
presented by us in [7], providing a unified pic-
ture of the variability observed in GRS 1915. In
[7], we modeled the large amplitude changes as
emptying and replenishing of the inner accretion
disk caused by a viscous-thermal instability. The
small radius observed during the quiescent pe-
riod was identified with the innermost stable orbit
3
Figure 1. Upper panel: 2-40 keV PCA light curve. Time zero corresponds to 1997 June 18th, 14:36 UT.
Middle and lower panels: corresponding inner radius and temperatures (see text)
around the black hole, while the large radius dur-
ing the burst phase was the radius of the emptied
section of the disk. The smaller oscillations were
interpreted as failed attempts to empty the inner
disk. As it can be seen from Figures 1 and 2, from
this observation we find that all variations, from
major events like the ones described in [7] to small
oscillations observed at the end of a large event,
can be modeled in exactly the same fashion. In
this scenario, the "flare state" presented in [7] is
simply a sequence of small events following a big
one, similar to the small oscillations in Figure 1.
Both the spectral evolution and the duration of
the event are determined by one parameter only,
namely the radius of the missing inner section
of the accretion disk.
It is natural to imagine
that the bigger the missing inner section of the
disk, the longer it will take to re-fill it. Indeed, if
we identify the re-filling time for each event with
the time spent in 'quiescence', this is what is ob-
served (Figure 3). Following [7], we can naturally
associate the re-filling time of an event tq to the
viscous time scale of the radiation-pressure dom-
inated part of the accretion disk. This can be ex-
pressed as tvisc = R2/ν, where ν = αcSH. using
the expressions for the scale-height H and sound
speed cS found in [10], we obtain:
tq ∼ tvisc = 30α−1
2 M −1/2
1
R7/2
7
M −2
18 s
(1)
where α2 = α/0.01, R7 is the radius in units
of 107 cm, M1 the central object mass in solar
masses, and M18 is the accretion rate in units of
1018 g/s. Notice that even the largest radii de-
rived here are well within the radiation-pressure
dominated part of the disk (see Equation 2 in [7]).
The line in Figure 3 represents the best fit to the
data with a relation of the form tq ∝ R7/2. The
fit is excellent, with the exception of the point
4
corresponding to the longest event. The qualita-
tive agreement with the theoretical expectation
is striking, although by substituting the appro-
priate values for the mass and accretion rate we
find that our best fit predicts rather small values
of α2 (0.004 and 0.05 for the Schwarzschild and
extreme Kerr cases respectively). This indicates
a small viscosity in the disk, although we stress
that tvisc is only a time scale, so that additional
corrections might be necessary in order to allow
a precise quantitative comparison.
Figure 2. Color-color diagrams for events of dif-
ferent duration. The 'low' intervals are corre-
spond to the populated clouds of points on the
low part of the diagrams. The sequence 'low'-
'high'-'low' moves clockwise
Let us follow an event and describe it in terms
M0 provided
of the model. The accretion rate
by the secondary is constant (within one observa-
tion). At the start of a quiescent period, the disk
has a central hole, whose radius is Rmax. Outside
M0 lies on a stable branch in the M − Σ
the hole,
M0
curve (see [7]). At all radii inside the hole,
lies on the unstable branch: the disk lies on the
lower (stable) branch. This means that the hole
does not have to be empty, but rather filled with
gas whose radiation is too soft to be detected.
Slowly the hole in the disk is re-filled by a steady
accretion rate M0 from outside. Each annulus of
the disk will move along the lower branch of its
S-curve in the M − Σ plane trying to reach M0.
The local density at each radius increases as the
annulus moves towards the unstable point at a
speed determined by the local viscous time scale.
During this period, no changes are observed in
the radius of the hole, since all the matter inside
does not radiate in the PCA band. In the XTE
data the disk appears to have a roughly constant
radius. The observed accretion rate during this
M0 since it is determined from the spec-
phase is
trum of the radiation coming from stable regions.
At the end of this phase, one of the annuli will
reach the unstable point and switch to the high-
M state, where the accretion rate is larger than
M0, causing a chain-reaction that will "switch
on" the inner disk. The observed accretion rate
M0 since
is now higher than the external value
it is determined from the spectrum of radiation
originating from the unstable part of the disk,
M is flowing. A smaller,
through which a higher
hot radius is now observed. At the end of the out-
burst, the inner disk runs out of fuel and switches
off, either jumping back to the M < M0 state or
possibly emptying completely. A new hole in the
disk is formed and a new cycle starts. Notice that
in this scenario the more "normal" state for the
source is the one at high count rates, where the
disk extends all the way to the innermost stable
orbit: in this state the energy spectrum is similar
to that of conventional black-hole candidates (see
[11]).
Not only the start and end of a major burst, but
also all the small amplitude oscillations within a
burst show the same timing signature of decaying
faster than they rise. This is in agreement with
what was already noticed in [7]: the rise time is
5
Figure 3. Correlation between the total length of an event and maximum inner radius of the disk. The
line is best fit with a power law with fixed index Γ=3.5.
determined by the speed at which a heating wave
moves through the central disk, while the faster
decay time is due to the rapid fall of matter into
the black hole (or ejection into a relativistic jet).
It has been found that when the hardness ratio
HR2 exceeds 0.1, the power density spectrum is
similar to the one observed in black-hole transient
systems during the Very High State (see [12]).
In these occasions, a strong 1-6 Hz QPO peak
was found, positively correlated with the count
rate [6]. The limit HR2>0.1 is an indication that
the source was in a quiescent state. Our spectral
results show that the fast timing features (both
QPO and band-limited noise, see [5]) cannot orig-
inate from the innermost regions of the optically
thick accretion disk, since those are missing dur-
ing the quiescent phases. The fact that the QPO
frequency increases with count rate is in quali-
tative agreement with the model, since a higher
count rate indicates a smaller inner disk radius,
and therefore shorter time scales. However, the
QPO seems to be associated to the power law
spectral component and not from the disk com-
ponent, making this phenomenology very difficult
to understand.
The radii for the disappearing region of the disk
found here are considerably smaller than that re-
ported in [7] from an observation where the length
of the events was substantially longer. This is en-
tirely due to the improved knowledge of the spec-
tral response of the PCA. Interestingly,the length
of the quiescent period and the maximum inner
radius observed in [7] are in agreement with the
curve in Figure 3 when the time is properly re-
normalized to take into account the difference in
accretion rate (see Equation 1).
Our model reduces the complication of the
spectral evolution of GRS 1915+105 to one pa-
rameter: the radius of the missing section of the
accretion disk. The structure of the variability
is however yet to be explained. The question to
be answered is: what determines the length of
REFERENCES
1. Mirabel I.F., & Rodr´ıguez L.F., 1994, Nature,
371, 46.
2. Bailyn C.D., et al., 1995, Nature, 378, 157.
3. Castro-Tirado A.J., Brandt S., & Lund S.,
1992, IAUC 5590.
4. Greiner J., Morgan E., & Remillard R.A.,
1996, ApJ, 473, L107.
5. Morgan E., Remillard R.A., & Greiner J.,
1997, ApJ, 482, 993.
6. Chen X., Swank J.H., & Taam R.E., 1997,
ApJ, 477, L41.
7. Belloni T., et al., 1997a, ApJ, 479, L145.
8. Belloni T., et al., 1997b, ApJ, 488, L109.
9. Taam R.E, Chen X., & Swank J.H., 1997,
ApJ, 485, L83.
10. Frank J., King A., & Raine D., 1992, "Ac-
cretion power in Astrophysics", Cambridge
Univ. Press, Cambridge.
11. Tanaka Y., & Lewin W.H.G., 1995, in "X-ray
binaries", eds. Lewin W.H.G., Van Paradijs
J., & Van den Heuvel E.P.J., Cambridge
Univ. Press, Cambridge.
12. Van der Klis M., 1995, in "X-ray binaries",
eds. Lewin W.H.G., Van Paradijs J., & Van
den Heuvel E.P.J., Cambridge Univ. Press,
Cambridge.
13. Pooley G.G., & Fender R.P., 1997, MNRAS,
292, 925.
14. Eikenberry S.S., et al., 1998, ApJ, in press.
15. Mirabel I.F., et al., 1998, A&A, 330, L9.
6
the next outburst? The model outlined above
does not provide an answer, but allows us to re-
formulate the question in more physical terms.
What determines how large the next missing sec-
tion of the disk will be? In some observations
the events are very regular, in some others they
seem to be random, and in some others no events
are observed at all. The latter might be part of
extremely long quiescent intervals. In the frame-
work of this model, it is clear that a regular quasi-
periodic structure of events arises naturally if the
radii of the holes are all similar. In the observa-
tion reported here a striking one-to-one relation
between quiescent and burst time is observed [8],
a relation which applies to the burst and i qui-
escent states of the observation presented in [7]
but is obviously not satisfied in other observa-
tions (see e.g.
[9]) nor during the "flare" state
of [7]. Moreover, as already mentioned, a few
observations among the ones we analyzed does
not fit this pattern and requires a different inter-
pretation (1996 June 16th is a template). Both
the timing structure and the spectral evolution in
these observation is radically different from what
presented here. Nevertheless, our model provides
a satisfactory interpretation of the cause of the
changes in the X-ray emission.
GRS 1915+105 is a remarkable X-ray source.
Despite it uniqueness (and because of it), we have
the chance to learn something fundamental about
accretion disks around black holes. Some of its
characteristics (the X-ray variability being the
major one) are indeed peculiar, but others (the
band-limited noise and QPO in power spectra,
the thermal disk and the hard power law com-
ponents in the energy spectra) are remarkably
similar to many 'normal' black hole candidates.
Whatever the different states in these sources are,
so far we have been observing them on long time
scales (years for persistent sources like Cyg X --
1 and GX 339 -- 4, months for black hole tran-
sients). Here we can observe more spectral and
timing variations within one hour of observation.
In addition, coordinated observations in different
bands are beginning to provide us with the first
links between what happens in the accretion disk
(observed in X-rays) and the ejection of relativis-
tic jets (observed in IR and radio) [13 -- 15].
|
0802.1168 | 1 | 0802 | 2008-02-08T16:05:02 | INTEGRAL observations of Her X-1 | [
"astro-ph"
] | Aims: We investigate the X-ray spectral and timing properties of the accreting X-ray pulsar Her X-1 observed with the INTEGRAL satellite in July-August 2005. Methods: The data analyzed in this work cover a substantial part of one main-on state of the source. The short-time scale pulse period development is measured. X-ray pulse profiles for different energy ranges and time intervals are constructed. Pulse-averaged and pulse-phase resolved broad band X-ray spectra are studied. Spectral changes during X-ray dips are explored. Results: The X-ray pulse profiles are found to change significantly during the period of observations. For the first time a strong spinup is measured within one 35 d cycle. Spectral characteristics observed during the X-ray dips are consistent with their interpretaion as due to partial covering as has been reported by several authors. The fundamental cyclotron absorption line is firmly observed in both pulse-averaged and pulse-phase resolved X-ray spectra. The energy, width, and the depth of the line are found to vary significantly with pulse phase. | astro-ph | astro-ph | Astronomy&Astrophysicsmanuscript no. herx1int170108r2
November 10, 2018
c(cid:13) ESO 2018
8
0
0
2
b
e
F
8
]
h
p
-
o
r
t
s
a
[
1
v
8
6
1
1
.
2
0
8
0
:
v
i
X
r
a
INTEGRAL observations of Her X-1
D. Klochkov1, R. Staubert1, K. Postnov3, N. Shakura3, A. Santangelo1, S. Tsygankov2, A. Lutovinov2,
I. Kreykenbohm1,4, and J. Wilms5
1 Institut fur Astronomie und Astrophysik, University of Tubingen, Sand 1, 72076 Tubingen, Germany
2 Space Research Institute, Profsoyuznaya str. 84/32, 117997 Moscow, Russia
3 Sternberg Astronomical Institute, Moscow University, 119992 Moscow, Russia
4 INTEGRAL Science Data Centre, Chemin d'Ecogia, 16, 1290, Versoix, Switzerland
5 Dr. Karl Remeis-Sternwarte, Astronomisches Institut, Universitat Erlangen-Nurnberg, Sternwartstr. 7, 96049 Bamberg, Germany
Accepted 29/01/2008
ABSTRACT
Aims. We investigate the X-ray spectral and timing properties of the accreting X-ray pulsar Her X-1 observed with the INTEGRAL
satellite in July -- August 2005.
Methods. The data analyzed in this work cover a substantial part of one main-on state of the source. The short-time scale pulse
period development is measured. X-ray pulse profiles for different energy ranges and time intervals are constructed. Pulse-averaged
and pulse-phase resolved broad band X-ray spectra are studied. Spectral changes during X-ray dips are explored.
Results. The X-ray pulse profiles are found to change significantly during the period of observations. For the first time a strong spin-
up is measured within one 35 d cycle. Spectral characteristics observed during the X-ray dips are consistent with their interpretaion as
due to partial covering as has been reported by several authors. The fundamental cyclotron absorption line is firmly observed in both
pulse-averaged and pulse-phase resolved X-ray spectra. The energy, width, and the depth of the line are found to vary significantly
with pulse phase.
Key words. X-ray binaries -- accretion disks -- neutron stars
1. Introduction
Discovered in 1972 by the Uhuru satellite (Giacconi et al. 1973;
Tananbaum et al. 1972) Her X-1 is one of the most intensively
studied accreting pulsars. Being part of a low mass X-ray bi-
nary it shows strong variability on very different time scales:
the 1.24 s spin period of the neutron star, the 1.7 d binary pe-
riod, the 35 d period of precession of the warped and tilted
accretion disk (Gerend & Boynton 1976; Shakura et al. 1999;
Ketsaris et al. 2000), and the 1.65 d period of the pre-eclipse dips
(Giacconi et al. 1973; Tananbaum et al. 1972). Due to the high
orbital inclination of the system (i > 80◦) the counter-orbitally
precessing warped accretion disk around the neutron star cov-
ers the X-ray source from the observer during a substantial part
of the 35 d period. This gives rise to the alternation of so-called
on (high X-ray flux) and off (low X-ray flux) states. The 35 d
cycle contains two on states -- the main-on and the short-on --
separated by ∼ 7 − 8 d off state. The X-ray flux in the mid-
dle of the main-on state is ∼ 4 − 5 times higher than that in
the middle of the short-on. The sharp transition from the off-
state to the main-on is called the turn-on of the source. Turn-
ons are usually used for counting the cycles. The 35 d period
manifests itself also in variations of the shape of X-ray pulse
profiles (Soong et al. 1990b; Trumper et al. 1986; Deeter et al.
1998; Scott et al. 2000) and modulation of optical light curves
(Gerend & Boynton 1976; Howarth & Wilson 1983).
In spite of a large amount of available observational data,
the physical interpretation of several observed phenomena in
Her X-1 are still unclear. For example, the physical mechanisms
responsible for the disk precession, the X-ray dips, and the evo-
lution of X-ray pulse profiles are highly debated. In this work
we present X-ray observations of Her X-1 performed by the
INTEGRALobservatory.
In Sect. 2 we describe the observations and data process-
ing. In Sect. 3.1 we present energy- and time-dependent pulse
profiles. In Sect. 3.2 we explore the behavior of the pulse pe-
riod. The spectral analysis is presented in Sects. 4.1 (pulse-
averaged spectra), 4.2 (spectral changes during X-ray dips),
and 4.3 (pulse-resolved analysis). The results are discussed in
Sect. 5. In Sect. 6 we present a summary and conclusions.
2. Observations and data processing
Her X-1 was observed by INTEGRAL (Winkler et al. 2003) on
July 22 -- August 3, 2005 (MJD: 53573 -- 53585 1). The obser-
vations were spread over ∼5 orbital revolutions of the satel-
lite: revs. 338, 339, 340, 341, and 342. In our analysis we used
data obtained with the instruments JEM-X (Lund et al. 2003),
IBIS/ISGRI (Lebrun et al. 2003; Ubertini et al. 2003), and SPI
(Vedrenne et al. 2003). We used the energy range ∼3 -- 20 keV
for JEM-X and ∼20 -- 100 keV for ISGRI and SPI (the spectrum
of Her X-1 is falling off steeply with energy so that there is no
signal above 100 keV). The data processing was performed with
the version 6.0 Offline Science Analysis (OSA) software dis-
tributed by ISDC (Courvoisier et al. 2003). An additional gain
correction based on the analysis of the position of the tungsten
background line was added (similar to Tsygankov et al. 2006).
For the pulse-phase resolved spectroscopy we also used the soft-
ware developed at IASF, Palermo for IBIS/ISGRI2 (Mineo et al.
1 MJD = JD -- 2400000.5
2 http://www.ifc.inaf.it/ferrigno/INTEGRALsoftware.html
D. Klochkov et al.: INTEGRAL observations of Her X-1
2005 August
1
3
2
2005 July
22
23
24
25
26
27
28
29
30
31
ISGRI
20−100 keV
40
30
20
10
0
rev. 342
rev. 340
rev. 338
rev. 341
rev. 339
JEM−X
3−20 keV
30
25
20
15
10
5
0
53572 53574 53576 53578 53580 53582 53584 53586
Time MJD
2
s
/
s
t
c
s
/
s
t
c
Fig. 1. X-ray light curves of Her X-1 obtained with ISGRI (top
panel) and JEM-X(bottom panel). The solid curve in the bottom
panel shows the ASM RXTE light curve averaged over many
35 d cycles (Klochkov et al. 2006). The INTEGRAL revolution
numbers are indicated.
2006; Ferrigno et al. 2007). This software allows to construct
phase-energy matrices from which pulse-resolved spectra and
energy-resolved pulse profiles can easily be extracted. The re-
sults of the analysis of pulse phase averaged spectra were in-
dependently checked using the software package developed at
Space Research Institute, Moscow.
Figure 1 shows the light curves of Her X-1 obtained with
ISGRI (top) and JEM-X (bottom). The observations partially
cover a main-on state of the source, 35 d phases φpre ∼0 -- 0.11
and φpre ∼0.20 -- 0.28 (φpre = 0 is the phase of the turn-on of the
source). The turn-on occurred at MJD ∼ 53574.7, corresponding
to orbital phase φorb ∼ 0.7. The solid curve on top of the JEM-X
light curve is the ASM RXTE (Levine et al. 1996) light curve
averaged over many 35 d cycles and renormolized to match the
one from JEM-X. All JEM-X data from revolution 341 are re-
jected by the ISDC team because of a strong solar flare. It is
seen that the light curve is quite typical for Her X-1. Pre-eclipse
dips (decrease of the flux just before the eclipse) are clearly seen.
3. Timing analysis
3.1. Pulseprofiles
To construct the 1.24 s pulse profiles of the source all photon
arrival times were converted to the solar system barycenter and
corrected for binary motion. The orbital parameters (Table 1)
used for the binary correction are based on an updated ephemeris
arrived at by combining historical timing data with the timing
results of the latest RXTE observations of Her X-1 (the cor-
responding paper is being prepared). As a folding period for
constructing pulse profiles we used Pspin = 1.23775836(10) s
which was obtained with the standard epoch folding method
(Leahy et al. 1983).
It was found that the pulse profiles do not change signifi-
cantly during the start (revs. 339 and 340) and during the end
(rev. 341) of the main-on state. Therefore, we constructed the
Table 1. Orbital parameters of Her X-1 used to correct the arrival
times of photons for orbital motion in the binary.
Porb
T π
2
a sin i
= 1.700167233 d,
= 53571.982111 (MJD),
= 13.1831 light sec.
pulse profiles separately for these two intervals without fur-
ther splitting in time. We excluded the data falling inside X-
ray eclipses, pre-eclipse dips or located outside the main-on
state (during the off-state the data are very noisy and show only
marginal sine-like pulsations). The resulting energy-resolved
pulse profiles are shown in Fig. 2. It is seen that the profiles
change significantly from the start to the end of the main-
on state. The main peak (pulse phase ∼0.6 -- 0.8) is stronger
and narrower at the end of the main-on. The interpulse (pulse
phase ∼0.1 -- 0.2) almost disappears. This behavior is typical for
Her X-1: analogous pulse profiles were observed by RXTE at
different times (see e.g. Kuster et al. 2005). We also studied the
energy dependence of the pulse fraction F determined as
F =
Imax − Imin
Imax + Imin
,
(1)
where Imax and Imin are background-corrected count rates in the
maximum and minimum of the pulse profile respectively. It is
clearly seen (Fig. 3) that the pulse fraction increases with en-
ergy (there may be a saturation at energies above ∼30 keV, how-
ever the statistics is marginal). The structure of the profile is also
energy-dependent, going from a triple-peaked shape at lower en-
ergies to an almost single-peaked at higher energies. The main
peak becomes narrower toward higher energies. A deeper anal-
ysis of the pulse profiles including the modeling of their shape
(similar to that performed in Panchenko & Postnov 1994 but tak-
ing into account 35 d variation) is ongoing and will be presented
elsewhere.
3.2. Pulseperiodbehavior
To explore more precisely the intrinsic (not affected by the
orbital motion in the binary) spin period of the neutron
star Pspin during the INTEGRAL observations we performed
a phase-connection analysis similar to Ferrigno et al. (2007);
Deeter et al. (1981) using well defined average pulse profiles.
The length of the time intervals used to produce a single pulse
profile was 3 -- 5 INTEGRAL Science Windows (6 -- 10 ksec). This
allowed to produce pulse profiles of a quality allowing to mea-
sure their shift in pulse phase with respect to other pulse profiles
within ∼0.01 s uncertainties. Figure 4 shows the observed pulse
arrival times minus the calculated times (using a constant pulse
period). If we assign the same statistical uncertainty of 0.01s (de-
termined from the scattering of data points) to each data point
then the linear fit to the data corresponding to a constant pulse
period (dashed line) gives χ2
red = 1.9 for 23 d.o.f. while the
quadratic fit corresponding to the presence of a non-zero Pspin
results in χ2
red = 1.0 for 22 d.o.f. We conclude, therefore, that the
neutron star is spinning up with − Pspin = (5.8 ± 1.5) × 10−13s/s.
This value is significantly higher than the averaged spin-up trend
of Her X-1 which is ∼ 1.1 × 10−13 s/s. Around the average spin-
up strong variability of the pulse period from one main-on to the
next is known to exist (see e.g. Fig. 1 in Staubert et al. 2006).
D. Klochkov et al.: INTEGRAL observations of Her X-1
3
Beginning of the main−on
φ ∼0.0−0.11
pre
End of the main−on
φ ∼0.20−0.28
pre
s
/
s
t
c
15
10
5
0
15
10
5
0
40
20
0
20
10
0
15
10
5
0
0.0
3−10 keV
10−17 keV
17−26 keV
26−32 keV
32−100 keV
0.5
1.0
Pulse phase
1.5
2.0
No JEM−X data
due to a solar flare
17−26 keV
26−32 keV
32−100 keV
0.5
1.0
Pulse phase
1.5
2.0
60
40
20
0
20
10
0
15
10
5
0
0.0
Fig. 2. Energy-resolved X-ray pulse profiles of Her X-1 at the start of the main-on (left column) and at the end of the main-on (right
column).
4. Spectral analysis
The spectral analysis has been performed using the XSPEC
v.11.3.2l spectral fitting package (Arnaud 1996). For the start of
the main-on the data from all three X-ray instruments have been
used. The fit, however, is mainly driven by JEM-X (. 20 keV)
and ISGRI (& 20 keV) data. The spectra obtained with SPI
(& 20 keV) have much lower statistics. For the end of the main-
on only the higher energy part of the spectrum is available since
the JEM-Xdata could not be used due to a solar flare (see above).
Following the OSA User Manuals3, we added systematic errors
at a level of 2% in quadrature to all JEM-X and ISGRI spec-
tral points in order to account for uncertainties in the response
matrices of the respective instruments.
4.1. Pulse-averagedspectra
For constructing the pulse-averaged spectrum, we used the data
from revolutions 339 and 340 (the start of the main-on state, see
Fig. 1). During this interval the data from all three X-ray instru-
ments are available. Eclipses and dips, which are clearly seen
3 http://isdc.unige.ch/index.cgi?Support+documents
in the light curve (Fig. 1) and have sharp boundaries, were ex-
cluded from the analysis. To model the broad band spectral con-
tinuum we used a power law with the exponential cutoff model
highecut of XSPEC:
Icont ∝
E−Γ,
E−Γ · exp (cid:16)−
E−Ecutoff
Efold
(cid:17),
if E ≤ Ecutoff
if E > Ecutoff,
(2)
where E is the photon energy; Γ, Ecutoff, and Efold are model pa-
rameters. A Gaussian line is added to model the iron fluores-
cence line at ∼6.5 keV. The cyclotron absorption feature is mod-
eled by a multiplicative absorption line with a Gaussian optical
depth profile. So, the final spectral function I is the following:
I = Icont · exp
−
(E − Ecycl)2
2σ2
cycl
+
−τcycl exp
+ K exp
−
(E − EFe)2
2σ2
Fe
,
(3)
where Ecycl and σcycl are the centroid energy and width of the
cyclotron line, EFe and σFe are the energy and width of the iron
emission line. τcycl and K are the numerical constants describing
4
D. Klochkov et al.: INTEGRAL observations of Her X-1
n
o
i
t
c
a
r
f
e
s
l
u
P
0.9
0.8
0.7
0.6
0.5
0.4
0.3
10
Energy [keV]
100
Fig. 3. Pulse fraction of Her X-1 as a function of energy.
]
s
[
t
f
i
h
s
l
e
s
u
P
0.04
0.02
0.00
−0.02
−0.04
−0.06
53576
53578
53580
Time [MJD]
53582
Fig. 4. Observed pulse arrival times of Her X-1 minus calculated
times using a constant pulse period. The dashed line shows a
linear fit corresponding to a constant pulse period while the solid
line shows a quadratic fit corresponding to a Pspin = (−5.8±1.5)×
10−13s/s .
the strength of the cyclotron line and iron line, respectively. The
count rate spectrum is shown in Fig. 5. The middle panel shows
the residuals after fitting the spectrum with the continuum model
without the iron emission line and the cyclotron absorption line.
Systematic features around 6 keV and 40 keV are clearly seen.
The bottom panel shows the residuals of our final fit which in-
cludes the two lines.
To account for large systematic uncertainties in the nor-
malization of the instruments we introduced in our models a
free multiplicative factor for each instrument: FISGRI, FSPI and
FJEM−X (for JEM-X the factor was fixed to 1.0). The best-fit pa-
rameters with corresponding 1σ(68%)-uncertainties are listed in
Table 2. The position of the cyclotron line at 38.2+0.8
−0.6 keV is
consistent with that measured by RXTE at a similar time (see
Table 1 in Staubert et al. 2007b) and, therefore, supports the cor-
relation between the cyclotron line energy and the maximum
main-on flux found by Staubert et al. (2007b).
We have checked the presence and the centroid energy of the
cyclotron line using different continuum models (such as e.g.
Fermi-Dirac cutoff, Tanaka 1986) as well as using different line
profiles (e.g. Lorenzian profile). It was found that the presence
and the energy of the feature are independent of the choice of
Table 2. Best fit spectral parameters of Her X-1 for the observa-
tions of revs. 339 and 340. 1σ(68%)-uncertainties (χ2
min + 1) for
one parameter of interest are shown.
Parameter
Value
Γ
Ecutoff [keV]
Efold [keV]
EFe [keV]
σFe [keV]
Ecycl
σcycl
τcycl
FJEM-X
FISGRI
FSPI
χ2
red/d.o.f.
0.91 ± 0.01
25.5+0.2
−0.3
9.0+0.3
−0.2
6.57+0.11
−0.13
0.50+0.17
−0.16
38.2+0.8
−0.6
9.3+0.4
−0.5
0.63+0.04
−0.06
1.0 (fixed)
0.81+0.01
−0.02
0.96 ± 0.02
1.19/207
the continuum and the line profile model. The highecut model
was found to provide a better description of the spectral con-
tinuum in Her X-1 with respect to other models. Additionally,
it allows to compare our results with previous observations of
the source, most of which were analyzed using the highecut
model (see e.g. Gruber et al. 2001). A discontinuity of the first
derivative of the spectral function at Ecutoff ∼ 25 keV (which can
be noticed in the residuals in Fig. 5) is far from the cyclotron
line energy and does not affect the line parameters (see however
Kretschmar et al. 1997).
We also explored the pulse-averaged spectrum obtained dur-
ing revolution 341 (end of the main-on). The continuum spectral
parameters are poorly constrained in this case due to the absence
of the low-energy (. 20 keV) part of the spectrum. However, the
found position of the cyclotron line Ecycl = 37.3+1.3
−1.1 keV is con-
sistent with that measured in revolutions 339 and 340.
4.2. X-raydips
In the X-ray light curve taken with INTEGRAL one can see
pre-eclipse dips on three subsequent orbits after the turn-on
(Fig. 1). The turn-on of the source can be considered as egress
from an anomalous dip at φorb ∼ 0.5 (according to the model
of Shakura et al. (1999) the source is partially screened by the
outer accretion disk rim in both cases: during the turn-on of the
source and during anomalous dips). Thus, the light curve ob-
tained with INTEGRAL contains three pre-eclipse dips and one
anomalous dip. Using the JEM-X data we calculated the ratio
H = I8−20 keV/I3−8 keV, where I8−20 keV and I3−8 keV are the count
rates in the respective energy ranges. The ratio H as a function
of time is shown in the bottom panel of Fig. 6. It can be seen
that H increases during X-ray dips most probably indicating the
low-energy absorption by the cold matter in the accretion stream
(during the pre-eclipse dips) and in the outer rim of the disk (dur-
ing the anomalous dip). Even though the result is dominated by
the last two pre-eclipse dips, the data of the first pre-eclipse dip
and the anomalous dip are consistent.
To check the absorption hypothesis we constructed JEM-X
and ISGRI spectra corresponding to the four dips and compared
them with those obtained outside the dips. We are aware of the
fact that due to different formation mechanisms of anomalous
and pre-eclipse dips their spectral characteristics might be dif-
V
e
k
/
c
e
s
/
s
t
n
u
o
c
1
1
.
0
1
0
.
0
3
−
0
1
4
−
0
1
5
−
0
1
5
0
5
−
5
0
5
−
5
10
Channel energy [keV]
20
50
Fig. 5. Top: The pulse-averaged X-ray spectrum of Her X-1 ob-
tained during revolutions 339 and 340 (start of the main-on);
middle: residuals after fitting the spectrum with the continuum
model described by Eq. 2 without the iron emission line and
cyclotron absorption line (to avoid confusion the SPI data are
not shown in this panel, their statistics is lower than that of
IBIS/ISGRI and JEM-X); bottom: the residuals after including
the iron and cyclotron lines in the model.
JEM−X 3−20 keV
25
20
15
10
s
/
s
t
c
5
0
6
5
4
3
2
1
0
V
e
k
8
−
3
I
/
V
e
k
0
2
−
8
I
53575
53576
53577
Time [MJD]
53578
53579
Fig. 6. The JEM-X 3 -- 20 keV light curve (top panel) and the ratio
of count rates in the harder (8 -- 20 keV) and softer (3 -- 8 keV)
energy ranges as a function of time (bottom panel).
D. Klochkov et al.: INTEGRAL observations of Her X-1
5
JEM−X
IBIS/ISGRI
both, absorbed and non-absorbed spectra to fit the data during
the dips. We used the same continuum model as in the previous
section. The final spectral model can be written as
SPI
is
the
cosmic
I(E) = A · [1 + α(E) · Icont],
(4)
where Icont is the continuum model described in the previous
section, and
α(E) = B · e−NHσbf (E),
photoabsorption
(5)
cross-section
where σbf(E)
per hydrogen atom for matter of
abundances
(Bałuci´nska-Church & McCammon 1992) used in the phabs
model of XSPEC and NH is the equivalent hydrogen column
density. A larger value of B implies a larger degree of ab-
sorbed flux. From the fit of the spectrum of the dips we found
B = 0.24 ± 0.03 and NH = (111+18
−19) × 1022 cm−2. The best-fit
parameters of the spectral continuum Icont are consistent with
those found from fitting the spectrum outside the dips (Table 2).
To compare this result with the spectrum outside the dips, the
latter was also fit by the partial covering model (Eq. 4). Both
spectra (inside and outside the dips) are shown in Fig. 7. For the
spectrum outside the dips (the upper one) the NH parameter was
found to be consistent with zero and the B factor, consequently,
was not restricted. We conclude therefore, that the X-ray data
of Her X-1 during the dips are consistent with the partial
absorption model.
4.3. Pulse-phase-resolvedspectra
It is well known that the X-ray spectrum of Her X-1 varies with
1.24 s pulse phase (Voges et al. 1982; Voges 1985; Soong et al.
1990a). So, we have performed a separate analysis of the spectra
accumulated in different pulse phase intervals. Figure 8 shows
an example of pulse resolved spectra of the source. Variability
of the continuum and the cyclotron line is clearly seen.
Since the shape of the pulse profile is changing significantly
from the start of the main-on (revs. 339, 340) to its end (rev. 341),
we analyzed the data from these two intervals separately. Phase
binning in each case was chosen to provide a similar statistics in
each spectral bin. Pulse phase zero is the same as in Fig. 2. The
X-ray spectrum of each phase bin was fitted with the spectral
model described in Sect. 4.1. Figures 9 and 10 show the best-fit
spectral parameters as a function of pulse phase for revs. 339 and
340 (left) and the rev. 341 (right). Vertical error bars correspond
to 1σ(68%)-uncertainties. The dotted line shows the correspond-
ing pulse profile. The spectral parameters in different pulse phase
bins are also listed in Table 3. Since no JEM-X data are available
for rev. 341 the power law photon index Γ could not be restricted
by the fit in this case. The iron line at ∼6.5 keV was detected only
in the phase bin 0.1 -- 0.54, corresponding to the off-pulse inter-
val. The data from other phase intervals do not require inclusion
of the line in the spectral model to obtain a good fit. For the data
corresponding to revs. 339 and 340, it is seen that all continuum
and the cyclotron line parameters are highly variable with pulse
phase. For revolution 341 the picture is not so clear due to the
absence of the low-energy (. 20 keV) part of the spectrum. Also
the total exposure time and the average flux of the source in rev.
341 are smaller than those during revolutions 339 and 340.
ferent. But due to the low statistics we cannot perform a separate
analysis of the two kinds of dips (the spectral parameters are not
restricted in this case). To fit the spectrum from the dips we used
the approach of Kuster et al. (2005). It was assumed that a com-
bination of direct and absorbed radiation is observed during the
dips. Thus, we used a partial covering model which combines
5. Discussion
5.1. Pulseprofilesandpulseperiod
In Sect. 3.1 we constructed 1.24 s X-ray pulse profiles of
Her X-1. The shape of the profiles is both energy- and time-
6
D. Klochkov et al.: INTEGRAL observations of Her X-1
Table 3. Best fit spectral parameters of Her X-1 in different pulse phase intervals. 1σ(68%)-uncertainties (χ2
of interest are shown.
min +1) for one parameter
Phase bin
Revs. 339 and 340
0.10 -- 0.54
0.54 -- 0.64
0.64 -- 0.70
0.70 -- 0.76
0.76 -- 0.85
0.85 -- 0.95
Rev. 341
0.00 -- 0.56
0.56 -- 0.64
0.64 -- 0.72
0.72 -- 0.79
0.79 -- 1.00
Γ
Ecutoff
Efold
Ecycl
σcycl
τcycl
0.92+0.02
−0.04
1.06+0.02
−0.02
0.73+0.02
−0.04
0.48+0.04
−0.02
0.79+0.04
−0.02
0.97+0.04
−0.09
--
--
--
--
--
25.8+0.7
−0.9
24.3+0.5
−0.3
24.3+0.2
−0.2
23.2+0.2
−0.3
25.2+0.8
−0.6
26.2+0.6
−0.6
22.1+0.3
−0.4
27.4+0.8
−0.6
25.2+0.2
−0.5
26.8+0.5
−0.6
23.8+0.4
−0.5
6.35+0.43
−0.20
10.11+0.45
−0.38
9.59+0.30
−0.31
9.48+0.25
−0.24
9.90+0.28
−0.29
7.86+0.47
−0.23
29.6+0.9
−0.5
39.6+1.1
−0.9
40.7+0.6
−0.5
40.8+0.4
−0.4
37.7+0.4
−0.5
32.9+1.3
−1.4
8.23+0.18
−0.11
11.45+0.58
−0.60
9.68+0.35
−0.37
9.54+0.28
−0.42
8.14+0.30
−0.35 --
33.5+0.1
−0.6
42.7+0.7
−0.8
39.5+0.3
−0.3
36.9+0.4
−0.2
6.86+0.93
−0.31
4.32+1.64
−1.09
4.84+0.72
−0.62
5.69+0.45
−0.48
5.30+0.74
−0.69
9.63+1.48
−1.01
0.18+0.60
−0.17
9.52+1.14
−0.65
5.67+0.25
−0.23
5.82+0.44
−0.44
--
0.67+0.11
−0.11
0.42+0.07
−0.06
0.66+0.06
−0.06
0.98+0.06
−0.06
0.71+0.06
−0.05
0.71+0.14
−0.10
≥ 1.0
1.39+0.13
−0.13
0.92+0.06
−0.07
1.04+0.07
−0.06
--
0.76−0.85
0.64−0.7
0.1−0.54
)
V
e
k
s
2
m
c
/
V
e
k
(
V
e
k
1
1
0
.
1
0
0
.
3
−
0
1
5
10
channel energy (keV)
20
50
Fig. 7. Examples of pulse-averaged spectra of Her X-1 from the
dips (lower curve) and outside the dips (upper curve). The dif-
ference of the spectral shape at lower energies caused by photo-
absorption inside the dips is clearly seen. The step-like feature
at ∼7 keV in the spectrum of the dips corresponds to the iron
K-edge (as implemented in the phabs model of XSPEC).
dependent. At higher energies the main peak becomes narrower
and the pulsed fraction increases (Figs. 2 and 3). Such a depen-
dence of the profile on energy is typical for accreting pulsars
(see e.g. Tsygankov et al. 2007). It can be understood in a sim-
ple purely geometrical picture: if the rotation axis of the neutron
star is inclined with respect to the axis of its magnetic field, it
is expected that the upper part of the accretion column emitting
softer photons is seen during a larger part of the neutron star spin
period while the emission region of harder photons, the "base"
of the accretion column, is screened by the neutron star surface
during most of the spin period. Additionally, the emission dia-
gram of harder photons most probably originating close to the
base of the accretion column is believed to be narrower than that
of softer photons due to the strong dependence of the scattering
cross-section on the angle between the direction of a photon and
the magnetic field lines (see e.g. Basko & Sunyaev 1976).
The difference of the pulse profiles in the left and right
columns of Fig. 2 (corresponding to the start and the end of
Fig. 8. Examples of pulse-resolved X-ray spectra of Her X-1
during revolutions 339 and 340. Pulse phase intervals are indi-
cated (the pulse phase zero is the same as in Fig. 2).
the main-on state, respectively) shows the evolution of the pro-
files with 35 d phase. Two models have been proposed to explain
these variations. Scott et al. (2000) suggested that the change
of the profile from the start to the end of the main-on state is
due to a "resolved occultation" of the emitting regions on the
neutron star by the precessing accretion disk. According to this
model, the disk progressively occults the neutron star towards
the end of the main-on. The other model assumes free preces-
sion of the neutron star (Trumper et al. 1986; Kahabka 1987,
1989; Prokhorov et al. 1990; Shakura et al. 1998). In this model
the observed behavior of the profile is explained assuming that
the emission region on the star surface has a complex shape
(due to the presence of higher multipole components of the neu-
tron star's magnetic field). Changing the viewing conditions of
the emitting region with the phase of the free precession causes
the observed variation of the pulse profile (Ketsaris et al. 2000;
Wilms et al. 2003). In pulse profiles obtained with INTEGRAL
one can see that the count rate at the maximum of the main peak
is higher at the end of the main-on than that during the start of
the main-on (Fig. 2). In the case of a progressive occultation one
would only expect a decrease in the intensity of any feature in
D. Klochkov et al.: INTEGRAL observations of Her X-1
7
the profile towards the end of the main-on. In the model of free
precession, however, viewing conditions of different parts of the
emission region may change with the precessional phase of the
star in such a way that the intensity of particular features of the
profile (e.g of the main peak) will increase towards the end of the
main-on. We argue, therefore, that the INTEGRALobservations
analyzed in this work support the model of a freely precessing
neutron star as an explanation of the time variation of the pulse
profile in Her X-1.
In Sect. 3.2 we determined the time derivative of the in-
trinsic (not affected by the orbital motion in the binary) pulse
period of the neutron star in Her X-1 during the INTEGRAL
observation. The pulsar was found to spin-up with the rate
− Pspin = (5.8±1.5)×10−13 s/s. This value is more than five times
larger than the mean spin-up trend of Her X-1 (∼ 1.1 × 10−13 s/s)
and consistent with the variations of Pspin between subsequent
35 d cycles, e.g. as seen by BATSE (Kunz 1996; Nagase 1989;
Sunyaev et al. 1988) and RXTE (Staubert et al. 2006). However,
this is the first time that such strong pulse period variations have
been measured within one 35 d cycle. A deeper analysis of these
pulse period variations and their corelation with the X-ray lu-
minosity (including the data from INTEGRAL presented here)
is currently being performed and will be presented in a separate
paper.
f
f
o
t
u
c
E
l
d
o
E
f
Γ
28
26
24
22
11
10
9
8
7
6
1.2
1.0
0.8
0.6
0.4
0.2
0.4
0.6
0.8
1.0
Pulse phase
5.2. AbsorptionduringX-raydips
0.0
0.2
0.4
0.6
0.8
1.0
Pulse phase
In Sect. 4.2 we have shown that the X-ray spectrum of Her X-1
obtained during X-ray dips can be modeled using a par-
tial covering model which assumes that the observed spec-
trum is a combination of direct and absorbed radiation. Such
a combination has already been observed in Her X-1 during
X-ray turn-ons (Davison & Fabian 1977; Becker et al. 1977;
Parmar et al. 1980; Kuster et al. 2005) and also during X-ray
dips (Vrtilek & Halpern 1985, although the observations in-
cluded only the lower energy part of the spectrum .20 keV).
According to the model of Shakura et al.
(1999) and
Klochkov et al. (2006) (which we adopt in this work), most of
the X-ray dips are produced by the occultation of the X-ray
source by the cold matter in the accretion stream which moves
out of the system's orbital plane. If the stream is not homoge-
neous but consists of "blobs" of material then the source will be
screened during some time intervals within a dip. Between these
intervals the source will be visible or seen through less dense
material. This picture seems to be confirmed by the complicated
behavior of the light curve inside the dips (Fig. 6). In our analy-
sis, in order to construct an X-ray spectrum of acceptable quality
we had to accumulate the flux during complete dips. The spec-
trum in this case will contain a superposition of the absorbed and
non-absorbed flux. The partial covering model fits the spectrum
of the dips clearly better than the simple absorption model.
5.3. Pulse-phasevariabilityoftheX-rayspectrum
The dependence of the spectral parameters in Her X-1 on pulse
phase observed by INTEGRAL basically confirms the results
obtained earlier with HEAO-1 (Soong et al. 1990b) but provide
better restriction of continuum parameters. The spectral varia-
tion with pulse phase is a common feature in X-ray pulsars (see
e.g. Kreykenbohm et al. 2004, and references therein) which is
usually attributed to the change of the viewing angle of the ac-
cretion region on the neutron star surface. Below we show that
Fig. 9. Best-fit spectral continuum parameters as functions of the
pulse phase for revolutions 339 and 340 (left) and the revolution
341 (right). The dashed curve shows the corresponding pulse
profile. Vertical error bars correspond to 1σ(68%)-uncertainties.
the pulse-phase variability observed in Her X-1 is basically con-
sistent with this interpretation.
As one can see in Fig. 9, the power law photon index Γ de-
creases during the main peak. It ranges from ∼1.0 in the off-
pulse to ∼0.5 close to the maximum of the main peak (see also
Lutovinov et al. 2000). This effect reflects the sharpening of the
main peak with energy which is observed in energy-resolved
pulse profiles constructed in Sect. 3.1 (Fig. 2). The spectral hard-
ening in the main peak can be explained by the dependence of
the optical depth on the angle between the line-of-sight and the
magnetic field lines (Pravdo et al. 1977). The closer the view-
ing direction is to the magnetic axis, the deeper we look into
the emission region where harder photons originate. Since it is
generally accepted that the main peak corresponds to the radia-
tion from one of the two magnetic poles, the viewing direction is
closest to the magnetic axis during the main peak which causes
the observed effect.
The exponential folding energy Efold is much higher during
the main peak than during the rest of the pulsation cycle. In
comptonized X-ray spectra (which are normally observed from
accreting pulsars) this parameter is believed to be proportional
to the plasma temperature (see e.g. Rybicki & Lightman 1979).
As already mentioned, during the main peak we see the emission
from higher optical depth. This means that the observed radia-
tion originates closer to the base of the accretion column where
the plasma temperature is higher.
The cyclotron line centroid energy Ecycl is also higher dur-
ing the main peak of the profile (Fig. 10). A relative amplitude
of its variation is ∼25%. As argued from Γ and Efold, during the
main peak we see X-rays that originate closer to the neutron star
8
D. Klochkov et al.: INTEGRAL observations of Her X-1
l
c
y
c
E
l
c
y
c
σ
l
c
y
c
τ
40
35
30
12
10
8
6
4
2
0
1.4
1.2
1.0
0.8
0.6
0.4
0.0
0.2
0.4
0.6
0.8
0.0
0.2
Pulse phase
0.4
0.6
0.8
1.0
Pulse phase
Fig. 10. The same as in Fig. 9 but for the cyclotron line parame-
ters (left: revolutions 339 and 340; right: revolution 341).
surface, i.e. where the magnetic field strength is higher. The in-
crease in Ecycl during the main peak is, therefore, qualitatively
consistent with this picture. If one assumes that the radiation
comes from a compact accretion column located at the magnetic
pole of the neutron star with a pure dipole field then the change
in the height of the observed emission region necessary to pro-
duce a ∼25% variation of the magnetic field strength is ∼1.1 km
(for a neutron star radius of 10 km). However, as it was shown by
Staubert et al. (2007a), the height of the emission and line form-
ing region in case of Her X-1 is most probably much smaller,
∼ 104 cm. Thus, a changing height of the emitting region above
the neutron star cannot explain the observed variability of the
cyclotron line energy with pulse phase in this source.
Another possibility for the line energy variability is to as-
sume a complicated structure of the magnetic field at the site
of X-ray emission. As shown by Shakura et al. (1991) and
Panchenko & Postnov (1994), the complex shape of the ob-
served pulse profiles in Her X-1 suggests that such a complicated
field structure including higher multipole components is indeed
present in the source. In this case, at different rotational phases
of the neutron star we will observe emission coming from the re-
gions corresponding to different sub-structures of the non-dipole
magnetic field with very different field strengths. Furthermore,
the numerical modeling of Her X-1 pulse profiles observed with
EXOSAT performed by Panchenko & Postnov (1994) suggests
that the magnetic field stength at the ring-like structure on the
neutron star's surface corresponding to the quadrupole compo-
nent of the field (which was introduced in the model) is indeed
by ∼27% lower than that at the magnetic pole. Thus, a compli-
cated magnetic field structure on the surface of the neutron star
can cause the observed 25%-variation of Ecycl with pulse phase.
6. Summary and conclusion
In this work we analyzed the INTEGRALobservations of the ac-
creting X-ray pulsar Her X-1. X-ray pulse profiles of the source
are constructed and their energy- and time-dependence are dis-
cussed. A strong spin-up is found during the INTEGRALobser-
vations. This is the first time that pulse period variation is mea-
sured within one main-on state of the source. The value of the
spin-up rate is consistent with the typical difference between the
values of the pulse period previously found in adjacent 35 d cy-
cles. Spectral changes during X-ray dips are studied. The X-ray
spectrum of the dips was modeled by a partial covering model
which assumes that the observed spectrum is a combination of
direct and absorbed radiation. This is in agreement with the
previously suggested interpretation assuming that the obscuring
matter (accretion stream) is not homogeneous but rather consists
of blobs of material (see e.g. Vrtilek & Halpern 1985). Energy-
resolved X-ray pulse profiles as well as the variation of cyclotron
line and continuum parameters with 1.24 s pulse phase were ex-
plored. The spectral changes with pulse phase are shown to be
qualitatively (for the cyclotron line energy -- also quantitatively)
consistent with those expected from the viewing conditions of
the complex emitting region which vary with the rotational phase
of the neutron star.
Acknowledgements. This research is based on observations with INTEGRAL,
an ESA project with instruments and science data centre funded by ESA
member states (especially the PI countries: Denmark, France, Germany, Italy,
Switzerland, Spain), Czech Republic and Poland, and with the participation of
Russia and the USA. The work was supported by the DFG grants Sta 173/31-2
and 436 RUS 113/717/0-1 and the corresponding RBFR grants RFFI-NNIO-03-
02-04003 and RFFI 06-02-16025, as well as DLR grant 50 0R 0302 and RFFI
07-02-01051. We also thank ISSI (Bern, Switzerland) for its hospitality during
the team meetings of our collaboration. AL acknowledges the financial support
of the Russian Science Support Foundation.
References
Arnaud, K. A. 1996, in Astronomical Society of the Pacific Conference Series,
Vol. 101, Astronomical Data Analysis Software and Systems V, ed. G. H.
Jacoby & J. Barnes, 17
Bałuci´nska-Church, M. & McCammon, D. 1992, ApJ, 400, 699
Basko, M. M. & Sunyaev, R. A. 1976, MNRAS, 175, 395
Becker, R. H., Boldt, E. A., Holt, S. S., et al. 1977, ApJ, 214, 879
Courvoisier, T. J.-L., Walter, R., Beckmann, V., et al. 2003, A&A, 411, L53
Davison, P. J. N. & Fabian, A. C. 1977, MNRAS, 178, 1P
Deeter, J. E., Pravdo, S. H., & Boynton, P. E. 1981, ApJ, 247, 1003
Deeter, J. E., Scott, D. M., Boynton, P. E., et al. 1998, ApJ, 502, 802
Ferrigno, C., Segreto, A., Santangelo, A., et al. 2007, A&A, 462, 995
Gerend, D. & Boynton, P. E. 1976, ApJ, 209, 562
Giacconi, R., Gursky, H., Kellogg, E., et al. 1973, ApJ, 184, 227
Gruber, D. E., Heindl, W. A., Rothschild, R. E., et al. 2001, ApJ, 562, 499
Howarth, I. D. & Wilson, B. 1983, MNRAS, 202, 347
Kahabka, P. 1987, NASA STI/Recon Technical Report N, 88, 19405
Kahabka, P. 1989, in ESA SP-296: Two Topics in X-Ray Astronomy, Volume 1:
X Ray Binaries. Volume 2: AGN and the X Ray Background, ed. J. Hunt &
B. Battrick, 447 -- 452
Ketsaris, N. A., Kuster, M., Postnov, K., et al. 2000, in Proc. Int. Workshop "Hot
Points in Astrophysics", JINR, Dubna, p. 192, ed. V. Belyaev, arXiv:astro-
ph/0010035
Klochkov, D. K., Shakura, N. I., Postnov, K. A., et al. 2006, Astronomy Letters,
32, 804
Kretschmar, P., Kreykenbohm, I., Wilms, J., et al. 1997,
in ESA Special
Publication, Vol. 382, The Transparent Universe, ed. C. Winkler, T. J.-L.
Courvoisier, & P. Durouchoux, 141
Kreykenbohm, I., Wilms, J., Coburn, W., et al. 2004, A&A, 427, 975
Kunz, M. 1996, PhD thesis, University of Tubingen, Germany
Kuster, M., Wilms, J., Staubert, R., et al. 2005, A&A, 443, 753
Leahy, D. A., Elsner, R. F., & Weisskopf, M. C. 1983, ApJ, 272, 256
Lebrun, F., Leray, J. P., Lavocat, P., et al. 2003, A&A, 411, L141
Levine, A. M., Bradt, H., Cui, W., et al. 1996, ApJ, 469, L33
Lund, N., Budtz-Jørgensen, C., Westergaard, N. J., et al. 2003, A&A, 411, L231
D. Klochkov et al.: INTEGRAL observations of Her X-1
9
Lutovinov, A. A., Grebenev, S. A., Pavlinsky, M. N., & Sunyaev, R. A. 2000,
Astronomy Letters, 26, 691
Mineo, T., Ferrigno, C., Foschini, L., et al. 2006, A&A, 450, 617
Nagase, F. 1989, PASJ, 41, 1
Panchenko, I. E. & Postnov, K. A. 1994, A&A, 286, 497
Parmar, A. N., Sanford, P. W., & Fabian, A. C. 1980, MNRAS, 192, 311
Pravdo, S. H., Boldt, E. A., Holt, S. S., & Serlemitsos, P. J. 1977, ApJ, 216, L23
Prokhorov, M. E., Shakura, N. I., & Postnov, K. A. 1990, Thirty-five-day cycle of
HER X-1: Synthesised optical light curves in the model of freely processing
neutron star, Tech. rep.
Rybicki, G. B. & Lightman, A. P. 1979, Radiative processes in astrophysics (New
York, Wiley-Interscience, 1979. 393 p.)
Scott, D. M., Leahy, D. A., & Wilson, R. B. 2000, ApJ, 539, 392
Shakura, N. I., Postnov, K. A., & Prokhorov, M. E. 1991, Soviet Astronomy
Letters, 17, 339
Shakura, N. I., Postnov, K. A., & Prokhorov, M. E. 1998, A&A, 331, L37
Shakura, N. I., Prokhorov, M. E., Postnov, K. A., & Ketsaris, N. A. 1999, A%A,
348, 917
Soong, Y., Gruber, D. E., Peterson, L. E., & Rothschild, R. E. 1990a, ApJ, 348,
641
Soong, Y., Gruber, D. E., Peterson, L. E., & Rothschild, R. E. 1990b, ApJ, 348,
634
Staubert, R., Klochkov, D., & Rodina, L. 2007a, A&A, [in preparation]
Staubert, R., Schandl, S., Klochkov, D., et al. 2006, in American Institute
of Physics Conference Series, Vol. 840, The Transient Milky Way: A
Perspective for MIRAX, ed. J. Braga, F. D'Amico, & R. E. Rothschild, 65 -- 70
Staubert, R., Shakura, N. I., Postnov, K., et al. 2007b, A&A, 465, L25
Sunyaev, R. A., Gilfanov, M. R., Churazov, E. M., et al. 1988, Soviet Astronomy
Letters, 14, 416
Tanaka, Y. 1986, in Lecture Notes in Physics, Berlin Springer Verlag, Vol. 255,
IAU Colloq. 89: Radiation Hydrodynamics in Stars and Compact Objects, ed.
D. Mihalas & K.-H. A. Winkler, 198
Tananbaum, H., Gursky, H., Kellogg, E. M., et al. 1972, ApJ, 174, L143
Trumper, J., Kahabka, P., Oegelman, H., Pietsch, W., & Voges, W. 1986, ApJ,
300, L63
Tsygankov, S. S., Lutovinov, A. A., Churazov, E. M., & Sunyaev, R. A. 2006,
MNRAS, 371, 19
Tsygankov, S. S., Lutovinov, A. A., Churazov, E. M., & Sunyaev, R. A. 2007,
Astronomy Letters, 33, 368
Ubertini, P., Lebrun, F., Di Cocco, G., et al. 2003, A&A, 411, L131
Vedrenne, G., Roques, J.-P., Schonfelder, V., et al. 2003, A&A, 411, L63
Voges, W. 1985, NASA STI/Recon Technical Report N, 85, 34112
Voges, W., Pietsch, W., Reppin, C., et al. 1982, ApJ, 263, 803
Vrtilek, S. D. & Halpern, J. P. 1985, ApJ, 296, 606
Wilms, J., Ketsaris, N. A., Postnov, K. A., et al. 2003, Izvestiya Akademii Nauk,
Ser. Fizicheskaya, 67, 310
Winkler, C., Courvoisier, T. J.-L., Di Cocco, G., et al. 2003, A&A, 411, L1
|
astro-ph/0401197 | 1 | 0401 | 2004-01-12T11:35:38 | Parameter properties and stellar population of the old open cluster NGC 3960 | [
"astro-ph"
] | We present a $BVI$ photometric and astrometric catalogue of the open cluster NGC 3960, down to limiting magnitude $V\sim22$, obtained from observations taken with the Wide Field Imager camera at the MPG/ESO 2.2 m Telescope at La Silla. The photometry of all the stars detected in our field of view has been used to estimate a map of the strong differential reddening affecting this area. Our results indicate that, within the region where the cluster dominates, the $E(V-I)$ values range from 0.21 up to 0.78, with $E(V-I)=0.36$ ($E(B-V)=0.29$) at the nominal cluster centroid position; color excesses $E(V-I)$ up to 1 mag have been measured in the external regions of the field of view where field stars dominate. The reddening corrected color-magnitude diagram (CMD) allows us to conclude that the cluster has an age between 0.9 and 1.4 Gyr and a distance modulus of $(V-M_V)_0=11.35$.
In order to minimize field star contamination, their number has been statistically subtracted based on the surface density map. The empirical cluster main sequence has been recovered in the $V$ vs. $V-I$ and in the $J$ vs. $J-K_S$ planes, using optical and infrared data, respectively. From these empirical cluster main sequences, two samples of candidate cluster members were derived in order to obtain the luminosity distributions as a function of the $V$ and $J$ magnitudes. The Luminosity Functions have been transformed into the corresponding Mass Functions; for $M>1 M_\odot$, the two distributions have been fitted with a power law of index $\alpha_V=2.95\pm0.53$ and $\alpha_J=2.81\pm0.84$ in $V$ and in $J$, respectively, while the Salpeter Mass Function in this notation has index $\alpha=2.35$. | astro-ph | astro-ph | Astronomy&Astrophysicsmanuscript no. n3960
(DOI: will be inserted by hand later)
November 3, 2018
4
0
0
2
n
a
J
2
1
1
v
7
9
1
1
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Parameter properties and stellar population of the old open
cluster NGC 3960⋆
L. Prisinzano1, G. Micela2, S. Sciortino2, F. Favata3
1 Dipartimento di Scienze Fisiche ed Astronomiche, Universit`a di Palermo, Piazza del Parlamento 1, I-90134 Palermo Italy
2 INAF - Osservatorio Astronomico di Palermo, Piazza del Parlamento 1, 90134 Palermo Italy
3 Astrophysics Division - Space Science Department of ESA, ESTEC, Postbus 299, 2200 AG Noordwijk, The Netherlands
Received 3 November 2003; accepted 18 December 2003
Abstract. We present a BVI photometric and astrometric catalogue of the open cluster NGC 3960, down to limiting magnitude
V ∼ 22, obtained from observations taken with the Wide Field Imager camera at the MPG/ESO 2.2 m Telescope at La Silla.
The photometry of all the stars detected in our field of view has been used to estimate a map of the strong differential reddening
affecting this area. Our results indicate that, within the region where the cluster dominates, the E(V − I) values range from
0.21 up to 0.78, with E(V − I) = 0.36 (E(B − V ) = 0.29) at the nominal cluster centroid position; color excesses E(V − I)
up to 1 mag have been measured in the external regions of the field of view where field stars dominate. The reddening corrected
color-magnitude diagram (CMD) allows us to conclude that the cluster has an age between 0.9 and 1.4 Gyr and a distance
modulus of (V − MV )0 = 11.35.
In order to minimize field star contamination, their number has been statistically subtracted based on the surface density map.
The empirical cluster main sequence has been recovered in the V vs. V − I and in the J vs. J − KS planes, using optical
and infrared data, respectively. From these empirical cluster main sequences, two samples of candidate cluster members were
derived in order to obtain the luminosity distributions as a function of the V and J magnitudes. The Luminosity Functions have
been transformed into the corresponding Mass Functions; for M > 1 M⊙, the two distributions have been fitted with a power
law of index αV = 2.95 ± 0.53 and αJ = 2.81 ± 0.84 in V and in J, respectively, while the Salpeter Mass Function in this
notation has index α = 2.35.
Key words. Open clusters - individual: NGC 3960 - photometry - astrometry - differential reddening - Luminosity and Mass
Function
1. Introduction
Old Galactic open clusters are stellar systems which can
survive only if they have a large initial population and lie
far from dense molecular clouds (van den Bergh & McClure,
1980; de La Fuente Marcos, 1997). For this reason they are less
numerous than the young open clusters and are mainly found in
the external region of the Galaxy, where there is a lower likeli-
hood of catastrophic encounters.
Cluster stars have the same distance and chemical compo-
sition; for this reason stellar clusters are among the most ap-
propriate objects to study the Initial Mass Function (IMF), that
is one of the most crucial ingredients for dynamical evolution
models and for galaxy formation and stellar evolution models.
Comparison of the IMF of old open clusters with the IMF of
nearby and young open clusters is fundamental to look for sim-
Send offprint requests to: [email protected]
⋆ Based on observations made with the European Southern
Observatory telescopes obtained from the ESO/ST-ECF Science
Archive Facility.
ilarities or discrepancies of the star formation processes in the
Galaxy disk, in different epochs and in different environments.
Accurate determination of the open cluster IMF is, how-
ever, a difficult task, because these objects are in general poorly
populated and, due to their location in the thin disk, are highly
contaminated by field stars seen along the same line of sight.
In the case of old open clusters, these difficulties are increased,
as they are very distant from the Sun and thus often strongly
obscured due to interstellar absorption. High-quality photome-
try and precise extinction determinations are crucial to obtain a
correct estimate of the IMF.
With an age of about 1.2 Gyr and a small spatial concentra-
tion, NGC 3960 is an interesting example of an old open clus-
ter suitable for IMF studies. It is a low-latitude cluster, located
at the celestial and Galactic coordinates RA=11h50m54s and
Dec=−55◦42′(J2000) and l =294◦.42 and b =+6◦.17, respec-
tively, about 1850 pc from the Sun.
Physical parameters of NGC 3960 were estimated by Janes
(1981), through photoelectric BV and intermediate-band David
Dunlop Observatory (DDO) photometry. One blue and one vi-
sual photographs were obtained in order to extend the mag-
2
Prisinzano et al.: NGC 3960
nitude sequences to fainter limits. From the BV photometry,
the author concludes that the cluster is (0.5 − 1.0) × 109 yr
old, while from the DDO photometry for several cluster giants
and using the Janes (1977) method, he derives the reddening
E(B − V ) = 0.29 ± 0.02. A distance modulus (V − MV )0 =
11.1 ± 0.2 and metallicity [Fe/H]=−0.30 have also been esti-
mated by the same author. [Fe/H]=−0.34 ± 0.08 was subse-
quently derived by Friel & Janes (1993) from a spectroscopic
study of 7 giant stars of NGC 3960.
Because of its low-latitude position, in the direction of
the Carina spiral feature (Miller, 1972), effects of differen-
tial reddening are expected across the region of the sky in-
cluding the cluster members. Feitzinger & Stuewe (1984) and
Hartley et al. (1986) indicate the presence in their catalogues
of two dark nebulae falling in the field of NGC 3960. More re-
cently, Dutra & Bica (2002) gave values E(B − V ) = 0.51
and E(B − V ) = 0.59 for these two overlapping dark nebu-
lae, respectively, derived from a 100 µm dust thermal emission
map. These regions are located about 13 arcmin from the clus-
ter center, where the E(B − V ) value is about 0.29 (Janes,
1981). This non negligible difference in the E(B − V ) value,
therefore, suggests the existence of variable extinction across
the field of NGC 3960.
In this work, using a procedure similar to that used by
Piotto et al. (1999) and von Braun & Mateo (2001), we attempt
to construct an extinction map of our field of view, in order to
correct the photometric catalogue of candidate cluster members
for differential reddening effects.
In Section 2, we describe the observations and the data re-
duction procedure used to obtain the photometric/astrometric
catalogue; in Section 3 we present the CMD and the method
used to obtain the extinction map and the reddening corrected
catalogue. In Section 4 we describe the method adopted to de-
rive the cluster parameters, while in Section 5 we describe the
statistical procedure used to recover the empirical cluster main
sequence from optical and infrared data and the photometric se-
lection of candidate cluster members. In Section 6, we present
the Luminosity Functions obtained from optical and infrared
data and the corresponding Mass Functions. Finally, we sum-
marize and discuss our results in Section 7.
2. Cluster BVI Photometry and Astrometry
2.1. Observations and Data Reduction
NGC 3960 is an open cluster observed in the Pre-FLAMES
Survey (PFS), carried out as part of the ESO Imaging Survey
(EIS), aimed to provide BVI imaging data for use in connection
with FLAMES, a multi-fiber spectrograph instrument on the
ESO VLT UT2 (KUEYEN) Telescope. The observations for
this survey were collected using the Wide Field Imager (WFI)
camera mounted at the Cassegrain focus of the MPG/ESO
2.2 m Telescope at La Silla (Chile). This instrument consists of
a 4 × 2 mosaic of 2k×4k CCD detectors with narrow inter-chip
gaps of width 23′′.8 and 14′′.3 along right ascension and decli-
nation, respectively, yielding a filling factor of 95.9%. With a
pixel size of 0.238′′, each chip covers a field of view of 8′.12 ×
16′.25, while the full field of view is 34′× 33′.
According to the PFS observing strategy, the images were
obtained in the BVI pass-bands; for each filter, one short ex-
posure of 30 seconds, useful to avoid saturating bright objects,
and two deep exposures, of 4 minutes each, have been taken.
In order to cover the inter-chip gaps, these observations were
dithered by 30′′ both in right ascension and declination. Table
1 gives the log-book of the observations and Figure 1 shows a
deep image of our field of view in the V band. The cluster is
approximately located in the central part of the field of view
where it is barely visible within a radius of about 7 arcmin.
Strong effects of differential reddening cause a non-uniform
spatial distribution of the field stars in the external regions of
the field of view. In addition to the science target exposures,
a set of technical frames for the instrumental calibration were
obtained during the same night.
To complement the optical data, we have utilized JHKS
infrared (IR) photometry from the All-Sky Point Source
Catalogue of the Two Micron All Sky Survey (2MASS)1
(Carpenter, 2001) available on the WEB2.
The instrumental calibration of the WFI CCD images has
been performed using the mscred package, a mosaic specific
task implemented as an IRAF package for the NOAO Mosaic
Data Handling System (MDHS).
The first stage of the data reduction process has been the
removal of the instrumental signatures from each raw CCD im-
age. First of all, we have subtracted the electronic bias using the
overscan region that has been subsequently removed. Exposure
bias patterns have been subtracted using a zero calibration mo-
saic obtained as the median of two bias frames. Finally, flat
fielding for each filter has been performed using a set of dome
flat fields combined into a median master flat. Due to some sat-
urated points, we could not use the sky flat fields.
Images in the I band have required a special treatment be-
cause of the strong effects of fringing. In order to remove this
instrumental artifact we have subtracted the fringing pattern
provided by the MPG/ESO 2.2 m Telescope team3, scaling it
to each exposures.
Instrumental magnitudes were obtained for each chip of
each mosaic image by using the DAOPHOT II/ALLSTAR
(Stetson, 1987) and ALLFRAME (Stetson, 1994) photomet-
ric routines. The point-spread functions (PSFs), determined for
each chip and filter, take into account the variation in the PSF
across the field of view of each WFI chip.
In order to obtain the aperture correction to the profile-
fitting photometry, growth curves were determined. First,
we have selected the stars used to define the PSF model;
next, all other objects were removed from the frames and
aperture photometry was carried out at a variety of radii.
DAOGROW (Stetson, 1990) was used to derive growth curves
and COLLECT to calculate the "aperture correction" coeffi-
1 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 Foundation
2 http://irsa.ipac.caltech.edu/
3 available at http://www.ls.eso.org/lasilla/Telescopes/2p2T/E2p2M/WFI
Table 1. Log-book of the observations.
Prisinzano et al.: NGC 3960
3
Target
RA (J2000) Dec (J2000)
Night
Filter
Exp. Time
EIS name
(h m s)
(d m s)
[s]
seeing
FWHM[′′]
NGC 3960 11 50 55.0 −55 41 36 24 -- 25 Feb 2000
(OC21)
B
V
I
1 × 30 + 2 × 240 0.99 -- 1.21
1 × 30 + 2 × 240 0.86 -- 1.38
1 × 30 + 2 × 240 0.81 -- 1.19
Fig. 1. A deep WFI image of the region around NGC 3960 taken in the B filter. The field center is RA=11h50m51s.6 and
Dec=−55◦42′07′′(J2000). Each chip is 8′.12 × 16′.25, while the full field of view is 34′× 33′. The cluster is approximately
located in the central part of the field of view where it is barely visible within a radius of about 7 arcmin. Strong effects of
differential reddening cause a non-uniform spatial distribution of the field stars in the external regions of the field of view.
cient for each chip from the difference between PSF and aper-
ture magnitudes of the selected stars.
Calibration to the Johnson-Cousin photometric system was
performed by means of a set of Landolt (1992) standard fields
(SA98, SA104, PG0918 and Ru 152), observed during the
same night through the 8 chips.
In order to derive the transformation coefficients to the stan-
dard system we have used equations of the form:
v = V + A0 + A1 × X + A2 × (B -- V),
b = B + B0 + B1 × X + B2 × (B -- V),
i = I + C0 + C1 × X + C2 × (V -- I).
where v, b and i are the standard star instrumental magnitudes,
corrected to the exposure time of 1 s; X is the airmass and V ,
(1)
4
Prisinzano et al.: NGC 3960
Table 2. Coefficients of the transformation to the standard system for each filter and for each chip. Col. 3 gives the number of
standard stars used to derive the zero points and the color terms (Col. 4 and 5), imposing the mean extinction coefficients of La
Silla (Col. 6); finally, Col. 7 gives the corresponding average standard errors.
Filter Chip # N. stars.
Zero Point
Color Term Extinction Av. St. Error
V
B
I
1
2
3
4
5
6
7
8
1
2
3
4
5
6
7
8
1
2
3
4
5
6
7
8
24
306
136
4
0
51
67
11
25
328
148
4
0
49
56
11
22
248
123
6
0
42
47
13
0.987 ± 0.007
1.022 ± 0.002
1.062 ± 0.003
1.095 ± 0.005
1.048 ± 0.003
0.982 ± 0.007
0.668 ± 0.012
0.723 ± 0.002
0.772 ± 0.002
0.810 ± 0.008
0.754 ± 0.004
0.697 ± 0.019
1.989 ± 0.012
2.042 ± 0.002
2.063 ± 0.003
2.072 ± 0.007
2.066 ± 0.005
2.009 ± 0.007
-0.097
-0.097
-0.097
-0.097
-0.097
-0.097
-0.097
-0.097
0.249
0.249
0.249
0.249
0.249
0.249
0.249
0.249
0.130
0.130
0.130
0.130
0.130
0.130
0.130
0.130
0.14
0.14
0.14
0.14
0.14
0.14
0.14
0.14
0.25
0.25
0.25
0.25
0.25
0.25
0.25
0.25
0.09
0.09
0.09
0.09
0.09
0.09
0.09
0.09
0.034
0.031
0.030
0.033
0.022
0.024
0.053
0.038
0.028
0.054
0.030
0.062
0.056
0.035
0.038
0.047
0.033
0.026
B and I are the magnitudes in the standard system. We first
attempted to derive all the coefficients (the zero points, the ex-
tinction and the color terms) for each chip, but by analyzing
the photometric residuals as a function of the time through the
night, we noticed that when the PG0918 standard field was ob-
served, the night was not photometric, as the mean residuals
systematically change with time. Since the problem does not
affect the other science observations, we have discarded all the
observations of the PG0918 field and we have only used the
Landolt fields SA98, SA104 and Rubin 152. In order to im-
prove the photometric calibration, the number of the standard
stars in each of these fields has been increased using secondary
standards defined in Stetson (2000). Unfortunately, neither the
Landolt's nor the Stetson's catalogue cover the chips 4 and 5 of
our field with a sufficient number of standards to determine in-
dependent coefficients for these chips. Neighboring chips have
been used to calibrate science observations in these chips, as
described at the end of this section. Imposing mean extinction
coefficients (Stetson, private communication), a zero point and
a color term were determined separately for the chips where a
sufficient number of standard observations were available. We
have found that the color terms computed for the chips 2, 3, 6
and 7 are consistent with each other, with variations . 0.01 in
B, . 0.02 in V and . 0.05 in I; for the chips 1 and 8 we have
found different values because they have been computed on a
smaller number of standard stars (see Table 2). Considering the
homogeneous characteristics of the instrument, we have chosen
to compute new zero points fixing the color term to the median
value of the color terms of the chips for which the number of
measured standard stars was sufficiently large. The resulting
coefficients are given in Table 2 together with the correspond-
ing average standard errors. Points above the 3 σ level were
not considered for the calibration. We note that the variations
among the photometric zero points of different chips (Table 2),
are . 0.11 in V , . 0.14 in B and . 0.08 in I, in agreement
with the results found in Zoccali et al. (2003).
As already mentioned, the NGC 3960 field observations
were dithered both in right ascension and declination to cover
the inter-chip gaps. This means that contiguous chips have sev-
eral stars in common. We have compared the magnitudes of
these stars to verify consistency of our photometry among dif-
ferent chips. The mean value and the root mean square of the
residuals for the stars with V < 20 are given in Table. 3.
Finally, the photometric zero points of the chips 3 and 6 have
been used to calibrate the stars in the chips 4 and 5, respec-
tively. Corrections . 0.02, . 0.01 and . 0.02 mag to these
photometric zero points have been found, for the magnitudes
V , B and I, respectively, using the photometric comparison of
the common stars in the overlapping regions.
Completeness of our star list and accuracy of our photom-
etry for each chip have been determined by tests with artificial
stars; a total of about 1350 stars have been added as described
in Prisinzano et al. (2003). Our data are 100% complete above
V = 20 and more than 50% complete above V ∼ 21, I ∼ 19.5
Table 3. The mean value and the root mean square of the V , B and I residuals of the common stars of contiguous chips.
Prisinzano et al.: NGC 3960
5
Chips < ∆V >
1-2
1-8
2-3
2-7
3-6
6-7
7-8
0.051
-0.006
0.005
0.021
0.046
-0.032
-0.065
rms
0.023
0.027
0.036
0.070
0.021
0.029
0.039
< ∆B >
0.043
0.033
0.004
0.043
0.035
-0.071
-0.015
rms
0.025
0.072
0.055
0.049
0.036
0.055
0.043
< ∆I >
0.054
0.009
-0.018
0.014
0.034
-0.023
-0.043
rms
0.047
0.020
0.031
0.051
0.017
0.026
0.027
Table 4. External photometric errors estimated from the artificial star experiments.
Chip
¯V
σext(V ) σext(B − V ) σext(V − I) Chip
1
1
1
1
1
2
2
2
2
2
3
3
3
3
3
4
4
4
4
4
16.721
17.617
18.663
19.746
20.878
16.738
17.916
18.856
19.989
21.185
16.755
18.064
19.313
20.321
21.318
16.980
17.910
18.942
19.987
21.277
0.003
0.006
0.010
0.027
0.065
0.003
0.006
0.015
0.028
0.094
0.003
0.007
0.019
0.034
0.093
0.003
0.006
0.016
0.031
0.074
0.005
0.009
0.021
0.047
0.150
0.006
0.009
0.021
0.062
0.188
0.006
0.015
0.033
0.086
0.227
0.004
0.013
0.034
0.087
0.231
0.004
0.007
0.013
0.032
0.083
0.004
0.004
0.010
0.027
0.093
0.003
0.009
0.019
0.032
0.126
0.003
0.007
0.012
0.033
0.065
5
5
5
5
5
6
6
6
6
6
7
7
7
7
7
8
8
8
8
8
¯V
16.873
18.024
19.128
20.111
21.456
16.911
18.049
19.051
20.079
21.459
16.995
18.084
19.144
20.155
21.441
16.804
17.897
18.983
20.088
21.283
σext(V ) σext(B − V ) σext(V − I)
0.003
0.007
0.016
0.033
0.113
0.003
0.006
0.015
0.040
0.110
0.003
0.009
0.016
0.035
0.120
0.003
0.007
0.013
0.033
0.095
0.006
0.013
0.037
0.062
0.213
0.006
0.013
0.034
0.085
0.254
0.007
0.018
0.037
0.070
0.246
0.005
0.013
0.036
0.072
0.170
0.003
0.007
0.018
0.043
0.105
0.004
0.006
0.018
0.043
0.116
0.003
0.010
0.013
0.039
0.095
0.003
0.007
0.013
0.038
0.081
and B ∼ 22. The data for the recovered artificial stars of each
chip have been sorted by the observed magnitude and divided
into 14 group of about 80 stars. For each group, the median
observed magnitude ¯V and the external errors have been com-
puted as in Stetson & Harris (1988). In Table 4, we report the
estimated errors for V > 16.5. The values for V < 16.5 are
less than 0.001 mag.
2.2. Astrometry
In order to determine the astrometric solution we have consid-
ered as reference catalogue the Guide Star Catalogue, Version
2.2.01 (GSC 2.2 STScI, 2001). We have used the Aladin Sky
Atlas to plot this catalogue on the finding chart of the open
cluster NGC 3960 field. For each chip we have chosen three
stars for which we have both the celestial coordinates of the
GSC 2.2 catalogue and the pixel coordinates. We have used
these stars as reference for the IRAF task ccxymatch; im-
posing a matching tolerance of 0′′.3, we have matched a total of
4224 stars. The IRAF task ccmap has been used to fit a trans-
formation between pixel coordinates and celestial coordinates
after applying a 3 σ clipping to the data. In our case, a tangent
plate projection with distortion polynomials has been chosen
for the sky projection geometry. The scatter plot and the distri-
butions of the RA and Dec residuals of the transformation are
shown in Figure 2 for the chips 2 and 3 where the cluster is
mainly concentrated. We note that the residual distribution is
highly concentrated with a Gaussian shape in both directions.
The mean offsets and the rms of the distributions, given on each
panel, show that the final accuracy is always better than 0′′.2.
In order to investigate eventual systematic dependence from the
position, we have plotted the RA and Dec residuals as a func-
tion of RA and Dec. The results for the chips 2 and 3 are shown
in Figure 3 where no noticeable systematic effects are evident.
Similar results have been found for the other chips.
3. The Color-Magnitude Diagrams
The color-magnitude diagrams of all the stars measured in
our field of view have been obtained using all sources with
SHARP parameter (Stetson, 1987) between −0.8 and 0.8. This
selection allows us to reject non-stellar objects such as semi-
resolved galaxies or blended double stars or cosmic rays. The
6
Prisinzano et al.: NGC 3960
Fig. 2. Scatter plot of the RA and Dec residuals of the transfor-
mation and distributions of ∆RA and ∆Dec obtained from the
comparison of our astrometric solution and the GSC 2.2 refer-
ence catalogue. The vertical and horizontal dashed lines mark
the mean residuals in RA and Dec. The results are shown only
for the chips 2 (left panel) and 3 (right panel), where the cluster
is mainly concentrated. The number of matched stars, the mean
values and the 1 σ rms in arcsec are given on the figures.
V vs. B − V and the V vs. V − I color-magnitude dia-
grams of the 39 411 selected stars are shown in Figure 4 (top
panels). Due to the Galactic position of the cluster, the dia-
grams are highly contaminated by foreground and background
stars making it difficult to discriminate the cluster main se-
quence. To facilitate the interpretation of the color-magnitude
diagrams, we have considered the distribution of the stars as a
function of distance from the cluster center. For this we have
first calculated the cluster centroid RAcen = 11h50m41s.8 and
Deccen = −55◦40′37′′.4 (J2000), as the median value of the
celestial positions of stars brighter than V=15.5. Therefore we
have performed stars counts by using our data to obtain a first
Fig. 3. Positional residuals, computed as in Figure 2, as a func-
tion of RA and Dec in the GSC 2.2 reference catalogue. The
results refer to chips 2 (left panel) and 3 (right panel), where
the cluster is mainly concentrated.
order approximation of the cluster size. We have derived the
surface stellar density by performing star counts in concen-
tric rings around the cluster centroid and then dividing them
by their respective areas. Figure 5 shows the resulting density
profile and the corresponding Poisson errors bars. We note a
flattening of the profile outside 7 arcmin. This density variation
suggests that most of the cluster members are located within 7
arcmin, whereas outside the Galactic disk population is dom-
inant. However, we consider this radius only an approximate
value where the cluster dominates the field. In fact, as expected
from dynamical evolution and mass segregation effects, faint
cluster stars can be located out of this radius. We note, how-
ever, that our cluster size estimate is significantly greater than
the value 2.75 arcmin given for the radius of this cluster by
Janes (1981), where only stars brighter than V = 17 were con-
sidered. From here on we refer to the above mentioned region
Prisinzano et al.: NGC 3960
7
inside the circle of 7 arcmin as the "cluster region" and to the
remaining part of our field of view as the "field region".
The color-magnitude diagrams obtained using all the stars
within the cluster region are shown in Figure 4 (bottom panels).
Both the diagrams exhibit the cluster main sequence down to
V = 17 but the faint star main sequence is still hidden by con-
taminating stars. The Turn Off (TO) point is roughly located at
(V = 14, B − V = 0.3), while the group of bright stars in the
red part of the CMD are clearly Red Giant Branch (RGB) stars
of NGC 3960.
Fig. 5. Star counts as a function of the distance from the cen-
troid for all stars detected in the field of NGC 3960.
3.1. The Differential Reddening Map
The cluster main sequence in Figure 4 appears rather broad in
both diagrams, with higher dispersion in the V vs. V − I CMD
(lower-right panel). Artificial-star tests indicate that our typ-
ical photometric errors for V < 17 are σ(B−V ) ∼ 0.007 and
σ(V −I) ∼ 0.004; therefore, after ruling out instrumental causes
and/or photometric errors as possible origin for this effect, we
have investigated a possible dependence of such broadening
from the positions of the stars in our field of view. In Fig 6 we
show the V − I vs V color-magnitude diagrams of 16 subre-
gions of about 8′.6×8′.3 in the NGC 3960 field. Celestial coordi-
nates corresponding to each subregion are given on the top and
right axes, respectively. The solid line on the color-magnitude
diagrams is the isochrone of 1.1 Gyr, metallicity Z = 0.01
computed by the Pietrinferni et al. (2003) (see Section 4).
We note that the cluster main sequence is dominant in the
central regions, while the number of stars in the upper main
sequence decreases in the external regions. Nevertheless, a dif-
ferent appearance of the main sequence as a function of the po-
sition is clearly evident as regards the theoretical model. This
effect could be due to the fact that the star cluster is present only
in one small well defined region; such a difference is however
also evident looking at the blue edge of the field population,
roughly located from 0.8 to 1.2 in (V − I). We note that this
shift does not depend from the chip to chip photometric zero
point offsets because it is present in different CMDs of stars
falling in the same chip; in addition differences in these CMD
are not consistent with photometric zero point residuals of ad-
jacent chips found in Section 2.1. Thus, we conclude that the
dependence of the appearance on the spatial positions of the
stars is real and can be due to differential reddening.
This conclusion is consistent with several previous studies
indicating the presence of dust clouds in the NGC 3960 field.
For instance, a region of 0.116 square degrees, located at the
RA=11h49m10s, Dec=−55◦31′40′′(J2000) and (l =294◦.15,
b =6◦.28), and falling in the north-west part of our field of view,
has been included in a catalogue of dark nebulae compiled by
Feitzinger & Stuewe (1984). In this study, many fields of the
ESO/SRC southern sky survey were examined looking for re-
gions of the sky where the apparent surface density is reduced
compared to the surrounding regions. The dark nebula falling in
our field of view was classified as a cloud with some structure
and relatively well defined edges. Another dark cloud, falling
also in the north-west part of our field of view, was identified
by Hartley et al. (1986) using the ESO/SERC Southern J sur-
vey. The catalogue obtained in this work includes a dark nebula
of minimum density of size 5′×3′ centered on RA=11h49m03s
and Dec=-55◦40′22′′(J2000) (l =294◦.16, b =6◦.13). The dis-
tances of these dark clouds are not known.
More recently, a value for the reddening of these re-
gions was given in the unified catalogue of dust clouds by
Dutra & Bica (2002). The reddening values were extracted
from the all-sky reddening E(B −V )FIR map of Schlegel et al.
(1998), based on a 100 µm dust thermal emission map with
resolution of ≃ 6′. For each region, the authors compare the
reddening in the nebula direction (E(B −V )cen) with the back-
ground reddening (E(B − V )bck) computed from the average
of the measured values in four background surrounding posi-
tions, used as reference. In our field they found E(B −V )cen =
0.51 and E(B − V )bck = 0.38, for the first dark nebula and
E(B − V )cen = 0.59 and E(B − V )bck = 0.55, for the
second one. These values are significantly different from the
cluster reddening value E(B − V ) = 0.29 estimated by Janes
(1981) as the average of reddening measurements ranging from
E(B − V ) = 0.20 and E(B − V ) = 0.34 obtained for 6 giants
in NGC 3960. Such reddening estimates indicate that strong
gradients of interstellar absorption are present in our field, con-
firming our conclusion on the presence of differential redden-
ing deduced by looking at the color-magnitude diagrams.
In order to quantify such effects on the color-magnitude
diagrams we have created two reddening maps, one for the stars
located at the cluster distance, using the cluster main sequence,
and a reddening map for the field star population, affected by
a further reddening spread due to distance. Both maps are used
to apply an average correction for differential reddening to the
photometric data.
8
Prisinzano et al.: NGC 3960
Fig. 4. The top panels show the V vs. B − V and the V vs. V − I color-magnitude diagrams of all the stars measured in our field
of view, while the bottom panels show the same diagrams of all the stars within the "cluster region".
As in Piotto et al. (1999), the main sequence reddening map
has been calculated as follows: we divided our field in 20 ×
20 subregions of about 1′.7×1′.7 where we have obtained the
CMDs. Such subregion size is a good compromise to have
a sufficient number of stars to identify the cluster main se-
quence, at least within 7 arcmin from the cluster centroid, and
high spatial resolution to map the differential reddening. We
have selected as fiducial region the subregion with coordinates
RA =[11h50m28s.3, 11h50m45s.6] and Dec =[-55◦41′13′′.2,
−55◦38′50′′.6] (J2000) where the cluster main sequence is well
defined and where our estimated cluster centroid is located; we
have fitted with a spline the points of the CMD where the star
Prisinzano et al.: NGC 3960
9
Fig. 6. V vs. V − I color-magnitude diagrams of 16 subregions of about 8′.6×8′.3 in the field of NGC 3960. The solid line is a
representative theoretical isochrone of 1.1 Gyr and metallicity Z = 0.01, calculated by Pietrinferni et al. (2003).
density is larger, defining the cluster main sequence; for each
star with 12 < V < 18.2 and 0.4 < (V − I) < 2.4, we have
calculated the distance from the fitted sequence along the red-
dening vector defined by the relation AV = 3.1/1.25 × E(V −
I), where E(V − I) = 1.25 × E(B − V ) (Munari & Carraro,
1996). The relative reddening has been calculated as the me-
dian value of the V − I component of these distances.
As in von Braun & Mateo (2001), we have obtained the
reddening map at the cluster distance shown in Figure 7, where
the darker regions correspond to the higher relative reddening;
for reference, the position of the fiducial region and the cluster
region are also indicated by the small box and the circle, re-
spectively. The reddening values in milli-magnitudes, relative
to the fiducial sequence and corresponding to each subregion,
are the top number in each pixel of the grid shown in Figure 8.
The bottom number in each pixel is the number of stars used to
determine the relative reddening in that pixel. The thicker box
indicates the position of the fiducial region. We note that higher
reddening regions correspond to lower star numbers with re-
spect to the average star number in the surrounding regions.
Nevertheless, we are aware that the relative reddening values
are significant for the central regions, where the cluster popula-
tion is dominant, but are not significant for the external regions
of our field of view where the stellar population is dominated
by field stars; in fact, these stars lay at different distances and
10
Prisinzano et al.: NGC 3960
Fig. 7. Greyscale reddening map of the NGC 3960 cluster members. In this figure, the darker regions correspond to higher relative
reddening, calculated with respect to the fiducial cluster main sequence. The position of the fiducial region and the cluster region
are also indicated as the small box and the circle, respectively.
the fiducial main sequence used to determine the reddening is
not representative in these regions.
In order to estimate the differential reddening affecting the
field stars, we have calculated the separation of each star with
18.5 < V < 22.0 and 0.0 < (V −I) < 2.8, from the blue edge
of the CMD in the fiducial region (von Hippel et al., 2002). The
relative reddening affecting these stars has been estimated as
the mode of the ∆(V − I) distribution. We note that we are ap-
proximating the relative reddening to a single value for all the
field stars, although we are aware that the relative reddening is
almost certainly spread out in distance. Using a single redden-
ing value for each subregion is an approximation, since clearly
we cannot determine the reddening for each star. However, our
value can be used to determine an average contamination of
cluster candidates, as will be discussed in Section 5.1. The ob-
tained results are shown in Figure 9, where the fiducial region
position is also indicated; the corresponding values in milli-
magnitudes and the number of the stars used to determine the
relative reddening are reported in Figure 10, where the posi-
tion of the fiducial region is marked by a thicker box. In Figure
9, the small circle and the circular line indicate the position
of the dark nebulae included in the Hartley et al. (1986) and
Feitzinger & Stuewe (1984) catalogues. As already mentioned,
Dutra & Bica (2002) estimated E(B − V ) = 0.59 in the re-
gion corresponding to the small circle. On the other hand, in
this region we have measured a shift of ∆E(V − I) = 0.45
with respect to the CMD blue edge in the fiducial region. Using
the relation given in Munari & Carraro (1996), this shift corre-
sponds to ∆E(B − V ) = 0.36. By subtracting this value to
the reddening estimated by Dutra & Bica (2002), we find, in
our fiducial region, E(B − V ) = 0.23, that is consistent with
the average value E(B − V ) = 0.29 given by Janes (1981) for
the central region of the cluster. This external test shows that
our differential reddening estimate is consistent with previous
reddening estimates.
We note that the two reddening maps are qualitatively con-
sistent with each other, both showing the two regions with high-
est reddening, located in the north-west and in the south-east
part of the field, and the lowest reddening region, located in
the north-east part of the field. In the cluster region the median
of the differences between the reddening values derived from
the two methods is AV = 0.08 mag with a standard devia-
tion of 0.25 mag. We have used the reddening values relative
to the cluster main sequence (Figure 8) to correct magnitude
and color measurements of the stars within the cluster region,
where we expect most of the cluster members to be located,
and the reddening values relative to the blue edge of the CMD
to correct magnitudes and colors of the stars in the field re-
gion, where the stellar population is dominated by field stars.
We find that, within the cluster region, the E(V − I) values
range from 0.21 up to 0.78, assuming that E(V − I) = 0.36
(E(B − V ) = 0.29 (Janes, 1981)) in the fiducial region; cor-
Prisinzano et al.: NGC 3960
11
Fig. 8. The top number in each pixel of the grid is the reddening values in milli-magnitudes, calculated with respect to the fiducial
main sequence and corresponding to each subregion, while the bottom number in each subregion is the number of stars used to
compute the relative reddening in that pixel. The fiducial region is marked by a thick box.
rections in E(V − I) up to 1 mag have been, instead, applied
for the most reddened field stars.
The reddening corrected V vs. V − I CMD in the cluster
region is shown in Figure 11 (left panel). We note that the main
sequence is better defined with respect to the uncorrected CMD
shown in the lower-right panel of Figure 4. The Turn Off and
the RGB stars, which were spread out in Figure 4, are now well
traced. Of course, our procedure computes an average correc-
tion on each pixel of our map and a residual spread in the CMD
remains.
dividual membership we can only constrain the age within the
range of ages of isochrones limiting most of the stars close to
the Turn Off. With this criterion the cluster age is between 0.9
and 1.4 Gyr.
Our results therefore suggest an age for NGC 3960 older
than that given in Janes (1981) and a distance of about 1850
pc, that is slight larger than the value given by the same author
(d=1660 pc). In the right panel of Figure 11 we show the V vs.
V − I diagram of the stars within the cluster region where the
adopted theoretical isochrones have been superimposed.
4. Cluster Fundamental Parameters
Our photometric data together with very recent stellar models
allow us to determine the cluster parameters, previously esti-
mated by Janes (1981) from a sample of only 318 stars with
V < 16.5.
In order to find the age and the distance of the cluster, we
have considered a set of different age theoretical isochrones re-
cently calculated by Pietrinferni et al. (2003), with metallicity
Z = 0.01, as spectroscopically determined by Friel & Janes
(1993). By fixing the reddening value E(B − V ) ≃ 0.29, de-
rived by Janes (1981), we have vertically shifted the set of red-
dened isochrones in the V vs. (V − I) CMD of the stars in
the cluster region. The distance that better fit the upper part of
main sequence stars (15 . V . 17) corresponds to a distance
modulus of (V − MV )0 = 11.35. Since we do not know the in-
5. Empirical Cluster Locus and Photometric
Selection of Candidate Cluster Members
5.1. Optical Data
As already mentioned, the reddening corrected V vs. V − I
CMD computed from the stars within the cluster region is
highly contaminated by field stars, as shown by the fact that
only the bright cluster main sequence is clearly visible. In or-
der to define a sample of candidate cluster members useful to
study the cluster population, we need to define photometrically
the complete cluster locus in the CMD. This result is usually
achieved by superimposing theoretical stellar models to the
observed CMD where the complete cluster main sequence is
clearly visible, which is not our case. To check the agreement of
theoretical isochrones, that suffer from a number of uncertain-
12
Prisinzano et al.: NGC 3960
Fig. 9. Greyscale reddening map of the field stars in the NGC 3960 field. In this figure, the darker regions correspond to the
higher relative reddening, calculated with respect to the blue edge of the CMD in the fiducial region. The position of the fiducial
region is indicated as the small box. The small circle and the circular line indicate the position of the dark nebulae included in
the Hartley et al. (1986) and Feitzinger & Stuewe (1984) catalogues.
ties (von Hippel et al., 2002; Grocholski & Sarajedini, 2003),
with our data, we have empirically recovered the cluster main
sequence using a statistical subtraction of the CMD density dis-
tribution, as described below. A similar approach has been used
by von Hippel et al. (2002).
We have used the field region to perform a statistical sub-
traction of the contaminating field stars in the CMD derived
in the cluster region. The result of this subtraction defines
the locus of the cluster main sequence (Chen et al., 1998;
von Hippel et al., 2002; Baume et al., 2003).
Using our reddening corrected photometric data, we have
defined a "clean field region" by selecting from the "field re-
gion" all those subregions in Figure 10 whose relative redden-
ing correction was between −0.150 and 0.200, corresponding
to the range of reddening, estimated for the field stars within
the "cluster region". We have chosen these limits in order to
consider only field stars affected by an absorption similar to
that of the cluster stars. For both the cluster and the field re-
gions, we have constructed a grid of boxes of ∆V = 0.125 and
∆(V − I) = 0.05 on the V vs. V − I plane. Thus, we have
built a greyscale CMD density map of the cluster by subtract-
ing the unreddened CMD density map of the field stars, from
the unreddened CMD density map computed for the "cluster
region". Figure 12 shows the resulting map where only the
boxes showing a 1 σ excess are shown. This excess should fol-
low the cluster main sequence. The cluster main sequence is
clearly present down to V ∼ 18, while the subtraction is nois-
ier for V > 18.5, the region of the cluster main sequence more
strongly contaminated by field stars, where the main sequence
exhibits a lower star density; as we have already mentioned, our
relative reddening correction suffers from the uncertainty due
to the spread of the reddening of individual stars, which pre-
vents us from obtaining an accurate field star CMD estimate.
This explains the lower star density in the low mass cluster
main sequence and the noise of the bluer field stars. However,
in this subtraction the faint cluster population (V > 18.5) is
well separated by the field star population this allowing us to
have a well defined empirical locus of the cluster main se-
quence.
In Figure 12,
the 0.9 and 1.4 Gyr
isochrones
(Pietrinferni et al., 2003) have been superimposed onto
the CMD density map using the cluster parameters found
in Section 4. The good agreement between the empirically
recovered cluster main sequence locus and the adopted theo-
retical model down to at least V ∼ 18, confirm both that the
parameters determined in Section 4 are suitable to the cluster
and that the adopted theoretical models reproduce the cluster
main sequence over the mass range for which the IMF will be
fitted.
Prisinzano et al.: NGC 3960
13
Fig. 10. The top number in each pixel of the grid is the reddening values in milli-magnitudes, calculated with respect to the blue
edge of the CMD in the fiducial region and corresponding to each subregion, while the bottom number in each subregion is the
number of stars used to compute the relative reddening in that pixel. The fiducial region is marked by a thick box.
We have used both the empirical cluster locus and the men-
tioned theoretical isochrones to define a photometric cluster
member sample. To this aim, we have defined a strip in the
observed CMD; the lower and upper limits take into account
the spread of the reddening determined above. Furthermore,
the upper limit of our strip is displaced upward by 0.75 mag
in order to include binary stars. We have defined as candidate
cluster members a total of 2119 stars lying in the cluster re-
gion that, within their photometric errors, belong to this strip.
The photometric/astrometric catalogue of these candidate clus-
ter members is given in Table 54, where we report RA and Dec
(J2000) coordinates in decimal degrees, an identification num-
ber for each star, the V , B, I magnitudes and their uncertain-
ties.
5.2. IR Data
As already stressed in the previous Sections, our field of view
is strongly affected by differential reddening, thus making un-
certain the field star CMD estimate and the study of the clus-
ter population. To have an external check of our results, we
have used near-infrared JHKs data from 2MASS survey of the
same field of view (Carpenter, 2001), since IR bands are not as
4 available in the electronic form at
the CDS via anony-
mous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via
http://cdsweb.u-strasbg.fr/cgi-bin
much sensitive to reddening effects as optical bands. To have
a sample as reliable as possible, only stars detected in all the
JHKS bands, with magnitudes measured from point spread-
function fitting or aperture photometry have been selected from
the 2MASS catalogue.
The first step has been to construct IR CMDs of the stars
in the cluster region. Figure 13 shows the J vs. J − KS CMD
of 1162 stars detected in this area; we can see that the cluster
main sequence is revealed from higher star density sequence
down to J ≃ 16.5, although it remains rather broad probably
due to a large scatter in the colors of the low Signal-to Noise
ratio sources.
In this diagram, the clump of stars at J ≃ 11 and (J −
KS) ≃ 0.7 are certainly Red Giant Branch stars of NGC 3960,
while the lower density stars in the almost vertical sequence
are mainly giant field stars. As already done with optical data,
we have built the greyscale density map of the cluster with IR
data by subtracting the field CMD density map from the J vs.
J − KS CMD density map computed for the "cluster region".
In order to estimate an optical-independent field star distribu-
tion, we have defined as "field region" the farther annulus area
with radius R between 30' and 45' from the cluster centroid.
Figure 14 shows the IR resulting map where only the boxes
having a 1 σ excess are shown. We note as the cluster main
sequence is well defined at least down to J = 15.8, that is the
2MASS completeness and reliability magnitude limit. In or-
14
Prisinzano et al.: NGC 3960
Fig. 11. Left panel: The V vs. V −I diagram of the stars within the cluster region, corrected for differential reddening. Right panel:
The V vs. V −I diagram of the stars within the cluster region, corrected for differential reddening, as compared to the Pietrinferni
et al. (2003) isochrones of age 0.9 and 1.4 Gyr, for the metallicity Z = 0.01. Using the color excess of E(B − V ) = 0.29 (Janes,
1981), a distance modulus of (V − MV )0 = 11.35 mag has been derived.
der to verify the cluster parameters, we have superimposed in
Figure 14 the 0.9 and 1.4 Gyr isochrones of Pietrinferni et al.
(2003), adopted also in the optical case, using the parameters
that we have determined in Section 4 and the reddening law
derived by Cardelli et al. (1989). We again find a good agree-
ment between theoretical models and observational data down
to J ≃ 15.5, while for fainter stars, the model do not repro-
duce the empirical cluster main sequence; this can be due both
to the strong uncertainties affecting IR color-temperature re-
lations (Grocholski & Sarajedini, 2003) and to the photomet-
ric errors in this magnitude range. However, the agreement of
the adopted theoretical model with the bright cluster main se-
quence confirms again the reliability of the cluster parameters
found in Section 4.
Once the empirical cluster main sequence has been recov-
ered from the IR data, we have applied, as done with the opti-
cal data, a photometric selection of candidate cluster members.
Using the adopted theoretical isochrone for the upper main se-
quence and a fiducial line for the lower one, we have defined
a strip in the CMD, including the boxes about the main se-
quence with a 1 σ excess; in this way we have selected a sam-
ple of 749 IR cluster candidates within the "cluster region" and
belonging, according to their photometric errors, to this strip.
The photometric/astrometric catalogue of these candidate clus-
ter members is given in Table 6 5, where we report RA and
Dec (J2000) coordinates in decimal degrees, an identification
number for each star, the V , B, I magnitudes and their uncer-
tainties.
6. Luminosity and Mass Functions
One of the main objectives of this study is to derive the stellar
mass function down to the limiting magnitude reached in this
survey, that is V ≃ 22.
In Section 5.1 we have found a sample of optical candi-
date cluster members, but such photometrically selected sam-
ple contains the contribution of foreground and background
sources that has to be subtracted in order to estimate the number
of stars belonging to NGC 3960. Since neither spectroscopic
observations, nor proper motions are available for this cluster,
we have used the following statistical approach to estimate the
number of the cluster members.
First, we have computed the total distribution of the optical
candidate members lying in the "cluster region", as a function
of the V magnitude; second, we have computed the V distribu-
tion of the candidate members falling in the "clean field region"
defined in Section 5.1, unreddened relatively to the fiducial re-
gion; then, this field distribution has been normalized to the
5 available in electronic form at
the CDS via anonymous
ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via
http://cdsweb.u-strasbg.fr
Prisinzano et al.: NGC 3960
15
Fig. 12. Greyscale CMD density map of the cluster obtained by
subtracting the V vs. V − I CMD density maps, computed for
the "clean field region" and for the "cluster region". Only boxes
with a 1 σ excess are shown. The solid and dashed lines are,
respectively, the 0.9 and 1.4 Gyr isochrones with metallicity
Z = 0.01, computed by Pietrinferni et al. (2003), scaled to the
observed CMD using the cluster parameters given in Section 4.
"cluster region" area and then subtracted to the total one. The
resulting V luminosity function of the cluster is shown in the
top panel of Figure 15, where the values of the absolute mag-
nitude (MV ), corresponding to the visual magnitudes are also
indicated in the top axis. We note that for V > 18.5, the distri-
bution is in some case consistent with the zero value and in one
case, is even negative(!). This result is consistent with the low
star density found in Figure 12 in the same magnitude range.
As already mentioned, this finding is very likely due to the crit-
ical field star distribution estimate about V = 19.5, where the
contribution of the field stars is dominant. For V > 20, the
deficiency of stars is probably due to incompleteness of our
catalogue that, based on the artificial star tests, is 100% com-
plete down to V ≃ 20. In addition, due to dynamical evolution,
a percentage of such faint cluster stars is expected to be found
also outside our estimated cluster region, that is where we have
estimated the field star distribution, and this makes it difficult
to have a correct estimate of the cluster luminosity function in
this magnitude range.
An independent IR luminosity distribution of the cluster
has been obtained using the same statistical approach and the
sample of IR candidate cluster members obtained in Section
5.2. As for the optical case, the contribution of field stars has
been computed using the "field region" defined in Section 5.2.
The resulting J luminosity function is shown in the bottom
panel of Figure 15. Also in this case, the result is reliable for
J < 15.8, that is in the magnitude range where the 2MASS
Fig. 13. J vs. J − KS CMD obtained from the 2MASS cata-
logue of the stars detected within 7 arcmin from the cluster cen-
troid. The diagram shows only stars revealed in all the JHKS
bands with magnitudes measured from point spread-function
fitting or aperture photometry.
catalogue is complete. This limiting magnitude corresponds to
V ∼ 17, for a main sequence star at the cluster distance.
In order to compare these results, obtained from two inde-
pendent surveys, we have transformed the V and J luminos-
ity functions into the corresponding mass functions. We have
used in both cases, the Mass-Luminosity Relation (MLR) com-
puted by Pietrinferni et al. (2003). The derived mass functions
are shown in Figure 16. As already stressed, the results are re-
liable for M > 1M⊙, where both surveys are complete. For
this mass range, we have computed a linear fit to both distri-
butions and have found that the two functions have a slope of
2.95 ± 0.53 and 2.81 ± 0.84, for the V and J luminosity func-
tions, respectively.
The agreement of the mass function obtained from the op-
tical data with the one obtained from IR data indicates that the
applied corrections for differential reddening and, therefore,
the field star distribution estimate, are reliable. Nevertheless,
the very large errors corresponding to the optical mass function
in the range M < 1M⊙ are the result of the already mentioned
problems in this mass range that prevent from having a reliable
estimate of the low mass function of NGC 3960.
7. Summary and Conclusions
This Chapter presents photometry and astrometry of the stars
in the field of the open cluster NGC 3960, falling in the 34 × 33
arcmin square field of the WFI camera of the MPG/ESO 2.2 m
Telescope. Our survey reaches a limiting magnitude V ∼ 22;
based on artificial star tests, it is 100% complete down to
16
Prisinzano et al.: NGC 3960
Fig. 14. Greyscale density map of the cluster obtained by sub-
tracting the J vs. J − KS CMD density map, computed for the
"field region" (cf Section 5.2), from the J vs. J − KS CMD
density map, computed for the "cluster region". Only boxes
with a 1 σ excess are shown. The solid and dashed lines are,
respectively, the 0.9 and 1.4 Gyr isochrones with metallicity
Z = 0.01, computed by Pietrinferni et al. (2003), scaled to the
observed CMD using the cluster parameters given in Section 4.
V ∼ 20, where the photometric accuracy is better than 6%.
Comparison of the positions of the stars in our survey with
those of the GSC 2.2, used as reference catalogue, indicates that
the final astrometric accuracy is <0′′.2.
The photometric data have been used to construct the V vs.
B − V and the V vs. V − I CMDs, where the high contribu-
tion of the disk population allows us to identify only the cluster
bright star main sequence, that appears rather broadened due to
strong differential reddening. A detailed spatial study of the V
vs. V − I CMD allows us to evaluate the relative reddening of
20 subregions, each of about 1′.7×1′.7, with respect to a fidu-
cial region corresponding to the cluster centroid. The relative
reddening has been computed from the cluster main sequence,
for the stars falling in the cluster region, and using the blue
edge of the CMD for the field stars, which, being at distances
different from that of the cluster, are affected by different red-
dening. Our results indicate that, within the cluster region, the
E(V − I) values range from 0.21 up to 0.78; corrections in
E(V − I) up to 1 mag have been, instead, applied for the most
reddened field stars.
The final reddening map traces regions of low star density,
that are consistent with two areas, indicated in literature as dark
nebulae, that are regions of the sky where the apparent surface
density of stars is reduced compared to surrounding regions.
Comparison of our reddening estimate in the smallest of these
dark nebulae, with the one reported in literature, allows us to
Fig. 15. Luminosity functions of NGC 3960, corrected for the
field star contribution, obtained from optical data (top panel)
and from IR data (bottom panel). The dashed lines indicate the
zero value of the distributions.
estimate a color excess E(B − V ) of about 0.23 in the region
corresponding to the cluster centroid. This value is very similar
to the value 0.29 given by Janes (1981) for the cluster.
Using the reddening corrected CMD and recent stellar
models, we have found that the cluster has an age between 0.9
and 1.4 Gyr, its distance modulus is (V − MV )0 = 11.35 and
it is located about 1850 pc from the Sun.
In order to minimize the field star contamination in the
CMD, a statistical subtraction of the CMD density map has
been performed to empirically recover the whole cluster main
sequence. Using the resulting density map, we have been
able to compare the empirical cluster main sequence with the
adopted theoretical models, thus allowing us to verify the re-
liability of our estimated cluster parameters. Furthermore, this
empirical cluster main sequence has been used to identify a
strip in the optical CMD where cluster stars are expected to be
located; 2119 candidate cluster members fall in this strip.
In order to have an external check of our results, we have
built an analogous CMD density map using IR data from the
2MASS catalogue. Taking advantage of the availability of the
entire sky coverage of the 2MASS catalogue and in order to
avoid the above mentioned dark nebulae, a different control re-
gion for field stars, located farther than the one used in the op-
tical case, has been used. The obtained empirical cluster main
sequence in the J vs. J − KS plane has been compared to the-
oretical models using the cluster parameters estimated in this
work. The result of this comparison shows, again, the correct-
ness of the estimated cluster parameters. As in the optical case,
the IR empirical main sequence has been used to define a strip
Prisinzano et al.: NGC 3960
17
pendent membership criterion and to determine individual red-
dening of each star.
Acknowledgements. We thank S. Cassisi and A. Pietrinferni for pro-
viding their most recent models and M. Zoccali and E. Flaccomio for
useful suggestions that greatly improved the presentation of this paper.
This work has been partially supported by MIUR.
References
Baume, G., V´azquez, R. A., Carraro, G., et al. 2003, A&A, 402,
549
Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345,
245
Carpenter, J. M. 2001, AJ, 121, 2851
Chen, B., Carraro, G., Torra, J., et al. 1998, A&A, 331, 916
de La Fuente Marcos, R. 1997, A&A, 322, 764
Dutra, C. M. & Bica, E. 2002, A&A, 383, 631
Feitzinger, J. V. & Stuewe, J. A. 1984, A&AS, 58, 365
Friel, E. D. & Janes, K. A. 1993, A&A, 267, 75
Grocholski, A. & Sarajedini, A. 2003, MNRAS, in press
Hartley, M., Tritton, S. B., Manchester, R. N., et al. 1986,
A&AS, 63, 27
Janes, K. A. 1977, PASP, 89, 576
Janes, K. A. 1981, AJ, 86, 1210
Kroupa, P. 2002, Science, 295, 82
Landolt, A. U. 1992, AJ, 104, 340
Miller, E. W. 1972, AJ, 77, 216
Munari, U. & Carraro, G. 1996, A&A, 314, 108
Pietrinferni, A., Cassisi, S., Salaris, A., et al. 2003, A&A, in
preparation
Piotto, G., Zoccali, M., King, I. R., et al. 1999, AJ, 118, 1727
Prisinzano, L., Micela, G., Sciortino, S., et al. 2003, A&A, 404,
927
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500,
525
Stetson, P. B. 1987, PASP, 99, 191
Stetson, P. B. 1990, PASP, 102, 932
Stetson, P. B. 1994, PASP, 106, 250
Stetson, P. B. 2000, PASP, 112, 925
Stetson, P. B. & Harris, W. E. 1988, AJ, 96, 909
van den Bergh, S. & McClure, R. D. 1980, A&A, 88, 360
von Braun, K. & Mateo, M. 2001, AJ, 121, 1522
von Hippel, T., Steinhauer, A., Sarajedini, A., et al. 2002, AJ,
124, 1555
Zoccali, M., Renzini, A., Ortolani, S., et al. 2003, A&A, 399,
931
Fig. 16. Comparison of the NGC 3960 mass function obtained
from the optical data (filled points) and from IR data (empty
triangles). The solid line is the power law fit to the mass func-
tion obtained from the V luminosity function, while the dashed
line is the power law fit to the mass function obtained from the
J luminosity function. ξ(M ) values are given in number per
logarithmic mass unit. The slopes of the power law fitting with
the rms residuals of the least squares fit, as obtained from the
two distributions, are also indicated.
in the J vs. J − KS CMD and therefore a sample of IR candi-
date cluster members.
From the two samples of candidate cluster members, ob-
tained from the optical and IR surveys, and using the respec-
tive control field, the luminosity distributions for the cluster,
as a function of the V and J magnitudes, have been com-
puted. Using the same MLR, the two distributions have been
transformed into two independent mass function determina-
tions of the cluster. In the mass range where the two data set
are complete (M < 1M⊙), the two mass functions have been
fitted to a power law having indices αV = 2.95 ± 0.53 and
αJ = 2.81±0.84 in V and in J, respectively, while the Salpeter
mass function in this notation has index α = 2.35. Our values
are both consistent with the data presented in the α vs. log M
plot of Kroupa (2002) for other open clusters.
The good agreement of the two distributions, furthermore,
suggests both that the reddening correction of our catalogue
and the estimates for the field star contamination are reliable
and correct. However, the low mass range is affected both by
completeness problems as well as by likely mass segregation
effects preventing a reliable derivation of the cluster luminosity
and mass functions for M < 1M⊙.
A future spectroscopic study of this region is planned in
order to measure the radial velocities crucial to have an inde-
|
astro-ph/0510833 | 1 | 0510 | 2005-10-31T20:18:02 | Results from the First INTEGRAL AGN Catalogue | [
"astro-ph"
] | We present results based on the first INTEGRAL AGN catalogue. The catalogue includes 42 AGN, of which 10 are Seyfert 1, 17 are Seyfert 2, and 9 are intermediate Seyfert 1.5. The fraction of blazars is rather small with 5 detected objects, and only one galaxy cluster and no star-burst galaxies have been detected so far. The sample consists of bright (fx > 5e-12 erg/cm**2/s), low luminosity (L = 2e43 erg/s), local (z = 0.020) AGN. Although the sample is not flux limited, we find a ratio of obscured to unobscured AGN of 1.5 - 2.0, consistent with luminosity dependent unified models for AGN. Only four Compton-thick AGN are found in the sample. This implies that the missing Compton-thick AGN needed to explain the cosmic hard X-ray background would have to have lower fluxes than discovered by INTEGRAL so far. | astro-ph | astro-ph |
RESULTS FROM THE FIRST INTEGRAL AGN CATALOGUE
Volker Beckmann1,2, Simona Soldi3,4, Chris R. Shrader1 and Neil Gehrels1
1NASA Goddard Space Flight Center, Exploration of the Universe Division, Code 661, Greenbelt, MD 20771, USA
2Joint Center for Astrophysics, Department of Physics, University of Maryland, Baltimore County, MD 21250, USA
3INTEGRAL Science Data Centre, 16 chemin d' ´Ecogia, 1290 Versoix, Switzerland
4Observatoire de Gen`eve, 51 chemin des Maillettes, 1290 Sauverny, Switzerland
ABSTRACT
We present results based on the first INTEGRAL AGN
catalogue. The catalogue includes 42 AGN, of which 10
are Seyfert 1, 17 are Seyfert 2, and 9 are intermediate
Seyfert 1.5. The fraction of blazars is rather small with 5
detected objects, and only one galaxy cluster and no star-
burst galaxies have been detected so far. The sample con-
sists of bright (fX > 5×10−12 erg cm−2 s−1), low lumi-
nosity ( ¯LX = 2 × 1043 erg s−1), local (¯z = 0.020) AGN.
Although the sample is not flux limited, we find a ratio
of obscured to unobscured AGN of 1.5 − 2.0, consistent
with luminosity dependent unified models for AGN. Only
four Compton-thick AGN are found in the sample. This
implies that the missing Compton-thick AGN needed to
explain the cosmic hard X-ray background would have to
have lower fluxes than discovered by INTEGRAL so far.
Key words: galaxies: active, catalogues, gamma rays:
observations, X-rays: galaxies, galaxies: Seyfert.
1.
INTRODUCTION
The X-ray sky as seen by satellite observations over the
past 40 years, shows a substantially different picture than
for example the optical band. While the visual night sky
is dominated by main sequence stars, Galactic binary sys-
tems and super nova remnants form the brightest objects
X-rays. Common to both regimes is the dominance of
active galactic nuclei (AGN) toward lower fluxes. In the
X-ray range itself, one observes a slightly different pop-
ulation of AGN at soft and at hard X-rays. Below 5 keV
the X-ray sky is dominated by AGN of the Seyfert 1 type;
above 5 keV the absorbed Seyfert 2 objects appear to be-
come more numerous. These type 2 AGN are also be-
lieved to be the main contributors to the cosmic X-ray
background above 5 keV (Setti & Woltjer 1989; Comas-
tri et al. 1995; Gilli et al. 2001), although only ∼ 50%
of the XRB above 8 keV can be resolved (Worsley et al.
2005).
The hard X-ray energy range is not currently accessible
to X-ray telescopes using grazing incidence mirror sys-
tems. Instead detectors without spatial resolution like the
PDS on BeppoSAX and OSSE on CGRO have been ap-
plied. A synopsis of these previous results is as follows:
the 2 -- 10 keV Seyfert 1 continua are approximated by a
Γ ≃ 1.9 powerlaw form (Zdziarski et al. 1995). A flat-
tening above ∼ 10 keV has been noted, and is commonly
attributed to Compton reflection (George & Fabian 1991).
There is a great deal of additional detail in this spectral
domain - "warm" absorption, multiple-velocity compo-
nent outflows, and relativistic line broadening - which are
beyond the scope of this paper. The Seyfert 2 objects are
more poorly categorized here, but the general belief is
that they are intrinsically equivalent to the Seyfert 1s, but
viewed through much larger absorption columns.
Above 20 keV the empirical picture is less clear. The
∼ 20 − 200 keV continuum shape of both Seyfert types is
consistent with a thermal Comptonization spectral form,
although in all but a few cases the data are not suffi-
ciently constraining to rule out a pure powerlaw form.
Nonetheless, the non-thermal scenarios with pure pow-
erlaw continua extending to ∼ MeV energies reported
in the pre-CGRO era are no longer widely believed, and
are likely a result of background systematics. However, a
detailed picture of the Comptonizing plasma - its spatial,
dynamical, and thermo-dynamic structure - is not known.
Among the critical determinations which INTEGRAL or
future hard X-ray instruments will hopefully provide are
the plasma temperature and optical depth (or Compton
"Y" parameter) for a large sample of objects.
The other major class of gamma-ray emitting AGN -
the blazars (FSRQs and BL Lac objects) are even more
poorly constrained in the INTEGRAL spectral domain
(for early INTEGRAL results see for example Pian et al.
2005).
Critical to each of these issues is the need to obtain im-
proved continuum measurements over the hard X-ray to
soft gamma-ray range for as large a sample of objects as
possible. INTEGRAL, since its launch in October 2002,
offers unprecedented > 20 keV collecting area and state
of the art detector electronics and background rejection
capabilities. Thus it offers hope of substantial gains in
our knowledge of the AGN phenomenon and in particu-
lar of the cosmic hard X-ray background.
The first INTEGRAL AGN catalogue offers the possibil-
ity to address these questions and to compare the results
with previous missions.
2. THE INTEGRAL AGN SAMPLE
Our INTEGRAL AGN sample1 consists of 42 extragalac-
tic objects, detected in the 20 − 40 keV energy band
with the imager IBIS/ISGRI. Spectra have been extracted
from IBIS/ISGRI, the spectrometer SPI, and the X-ray
monitor JEM-X in order to cover the energy range from
3 − 500 keV. The list of sources with their redshift, the
optical counterpart type, the flux in the 20 -- 40 keV band
as measured by ISGRI, the luminosity in the 20 -- 100
keV band, and the intrinsic absorption as measured at soft
X-rays is given in Table 1. Details on the analysis and
on individual spectra can be found in Beckmann et al.
(2006). The distribution of sources in the sky is shown in
Figure 2.
The Seyfert type AGN found in the sample are prefer-
entially low redshift objects. Figure 1 shows the distri-
bution of redshifts in the sample. The blazars all show
higher redshifts (0.15 < z < 2.51) and are not in-
cluded in the histogram. The one object on the right
is PG 1416 -- 129 (z = 0.1293) and is an anomalous ra-
dio quiet quasar with similar spectroscopic properties as
radio-loud sources (Sulentic et al. 2000).
In order to investigate the AGN subtypes, we have de-
rived averaged spectra of the Seyfert 1 and 2 types, as
well as for the intermediate Seyferts and the blazars.
The average Seyfert 1 spectrum was constructed using
the weighted mean of 10 ISGRI spectra, the Seyfert 2
composite spectrum includes 15 sources, and 8 objects
form the intermediate Seyfert 1.5 group. The two bright-
est sources, Cen A and NGC 4151, have been excluded
from the analysis as their high signal-to-noise ratio would
dominate the averaged spectra. The average spectra have
been constructed by computing the weighted mean of all
fit results on the individual sources. In order to do so,
all spectra had been fit by an absorbed single powerlaw
model. When computing the weighted average of the var-
ious sub-classes, the Seyfert 2 objects show flatter hard
X-ray spectra (Γ = 1.95 ± 0.01) than the Seyfert 1.5
(Γ = 2.10±0.02), and Seyfert 1 appear to have the steep-
est spectra (Γ = 2.11 ± 0.05) together with the blazars
(Γ = 2.07 ± 0.10).
1http://heasarcdev.gsfc.nasa.gov/docs/integral/spi/pages/agn.html
2
Figure 1. Redshift distribution of the AGN detected by
INTEGRAL. Blazars are not shown. The average red-
shift is ¯z = 0.020. The object on the right is the quasar
PG 1416 -- 129.
The Seyfert type classification of the objects is based on
optical observations. An approach to classifying sources
according to their properties in the X-rays can be done
by separating the sources with high intrinsic absorption
(NH > 1022 cm−2) from those objects which do not
show significant absorption. The distribution of absorp-
tion at soft X-rays for the AGN sample is shown in Fig-
ure 3. The black part of the histogram represents the
Seyfert 2 AGN (including the Seyfert 1.8 and Seyfert 1.9
subtypes).
It has to be pointed out that not all objects
which show high intrinsic absorption in the X-rays are
classified as Seyfert 2 galaxies in the optical, and the
same applies for the other AGN sub-types. Neverthe-
less a similar trend in the spectral slopes can be seen:
the 21 absorbed AGN show a flatter hard X-ray spec-
trum (Γ = 1.98 ± 0.01) than the 13 unabsorbed sources
(Γ = 2.08 ± 0.02). The blazars have again been excluded
from these samples.
Among the Seyfert 2 galaxies we find 4 objects with low
absorption (NH < 1023 cm−2), 7 with intermediate ab-
sorption (NH = 1023 − 1024 cm−2), and four Compton
thick AGN (NH > 1024 cm−2).
Although the INTEGRAL AGN sample discussed here
is not a complete flux limited one, the number counts
give a first impression regarding the flux distribution
within the sample (Fig. 4). Excluding the two bright-
est objects (Cen A and NGC 4151) and the objects with
fX < 2 × 10−11 erg cm−2 s−1 (where the number counts
shows a turnover), the number counts relation shows a
gradient of 1.4 ± 0.1, and is consistent with the value of
1.5 expected for Eucledian geometry and no evolution in
the local universe.
3
Figure 2. The distribution of INTEGRAL AGN in the sky in Galactic coordinates. Seyfert 1 are marked with up-triangles,
Seyfert 1.5 with down-triangles, Seyfert 2 with circles, blazars with squares, optically unidentified with asterisks, and the
Coma Cluster is represented by a star.
MCG +8−11−11
3C 111
NGC 4507
IC 4329A
PG 1416−129
3C 273
NGC 4388
NGC 4945
Circinus galaxy
NGC 4151
Cen A
Figure 3. Distribution of intrinsic absorption, as mea-
sured in the soft X-rays. The Seyfert 2 objects (including
the Seyfert 1.8 and 1.9 subtypes) are shown in black.
Figure 4. Number counts of the INTEGRAL AGN sample.
The brightest objects have been labelled.
4
X = 1.7 ± 0.4. The ratios change slightly when tak-
ing into account only those objects which belong to the
complete sample with an ISGRI significance of 7σ or
higher (Beckmann et al. 2006). This sub-sample includes
32 AGN, with 18 obscured and 10 unobscured objects
(absorption information is missing for the remaining four
objects). Using only the complete sample gives a simi-
lar ratio of X = 1.8 ± 0.5. Splitting this result up into
objects near the Galactic plane (b < 20◦) and off the
plane shows for all objects a ratio of X = 3.3 ± 1.1 and
X = 1.1 ± 0.5, respectively. This trend shows that the
harder spectra of those objects, where the absorption in
the line of sight through the Galaxy is low compared to
the intrinsic absorption, are more likely to shine through
the Galactic plane.
(1999) studied a large sample of Seyfert
Risaliti et al.
2 galaxies focusing especially on the intrinsic absorp-
tion measured at soft X-rays. They also find a frac-
tion of 75% of Seyfert 2 with an intrinsic absorption
NH > 1023 cm−2, but a 50% fraction of Compton-thick
objects with NH > 1024 cm−2, where the INTEGRAL
sample only finds 4 objects (27 %).
Optical studies in the local universe find evidence that
type 2 AGN are about a factor of four more numerous
than type 1 AGN (Setti & Woltjer 1989; Comastri et al.
1995).
In X-rays the situation is similar, although not
all Seyfert 1 objects show low intrinsic absorption and
vice versa (see Fig. 5). Recent studies have shown that
the fraction of absorbed sources depends both on lumi-
nosity and redshift in a way that the fraction of type 2
AGN increases towards higher redshifts and lower lumi-
nosity (e.g. Gilli et al. 2001; Ueda et al. 2003; La
Franca et al. 2005). Ueda et al.
(2003) sudied 247
AGNs in the 2 -- 10 keV band with luminosities in the
range LX = 1041.5 − 1046.5 erg s−1, similar to the lu-
minosity range of the INTEGRAL AGN but extending
to higher redshifts (z = 0 − 3) and to fainter fluxes
(fX = 10−10−3.8×10−15 erg cm−2 s−1). They find that
the number of AGN decreases with the intrinsic X-ray lu-
minosity of the AGN and therefore favour a luminosity
dependend density evolution (LDDE) to explain the lu-
minosity function in the 2 − 10 keV band. La Franca et
al. (2005) used an even larger sample and confirmed the
necessity of a LDDE model, where low luminosity AGN
peak at z ∼ 0.7, while high luminosity AGN peak at
z ∼ 2.0. In addition, they find evidence that the fraction
of absorbed (NH > 1022 cm−2) AGN decreases with the
intrinsic luminosity in the 2 − 10 keV energy range. Con-
sistent with our study, La Franca et al. also find a ratio of
X = 2.1 at LX = 1042.5 erg s−1.
All these results based on the 2 -- 10 keV band are con-
sistent with the findings of the INTEGRAL AGN sam-
ple at higher energies (20 − 40 keV).
It is surprising
though that different from the findings of Risaliti et al.
(1999) we do not detect a large fraction of Compton-
thick AGN. These AGN, if exisiting, could explain the
peak in the hard X-ray background around 30 keV (e.g.
Maiolino et al. 2003). In view of recent results this lack
of Compton-thick objects is explainable by the type of
Figure 5.
Intrinsic absorption, as measured at soft X-
rays, versus the luminosity in the 20 -- 100 keV band.
Circles represent Seyfert 2 type AGN (including Sy 1.8
and Sy 1.9), triangles other Seyfert types.
Figure 5 shows the intrinsic absorption of the objects, as
measured in the soft X-rays, versus the luminosity in the
20 -- 100 keV energy band as measured by INTEGRAL.
There is no discernable correlation between those values.
The optically classified Seyfert 2 objects naturally popu-
late the area with higher intrinsic absorption.
3. DISCUSSION
The typical INTEGRAL spectrum can be described by
a simple powerlaw model with average photon indices
ranging from Γ = 2.0 for obscured AGN to Γ = 2.1
for unabsorbed sources. The simple model does not give
reasonable results in high signal-to-noise cases, where
an appropriate fit requires additional features such as a
cut-off and a reflection component (Soldi et al. 2005,
Beckmann et al.
2005). The results presented here
show slightly steeper spectra than previous investigations
of AGN in comparable energy ranges. The same trend
is seen in the comparison of Crab observations, where
the INTEGRAL/ISGRI spectra also appear to be slightly
steeper than in previous observations, and in comparison
of RXTE and INTEGRAL/ISGRI spectra of Cen A (Roth-
schild et al. 2006). This trend might be based on im-
proper calibration and/or background subtraction.
The average properties, like spectral slope, redshift, lu-
minosity, are rather similar compared to previous studies
(e.g. Zdziarski et al. 1995, Gondek et al. 1996).
Comparing the ratio X of obscured (NH > 1022 cm−2)
to unobscured AGN we find in the INTEGRAL data that
AGN detected so far by INTEGRAL. All objects are local
AGN, with a mean redshift of ¯z = 0.020 (Fig. 1). The
objects are bright (f20−40 keV > 5 × 10−12; Fig. 4), but
have low luminosities ( ¯LX = 2 × 1043 erg s−1). This
still leaves room in the parameter space of AGN to locate
Compton-thick AGN. Those objects could have lower
fluxes and therefore even lower luminosities in the lo-
cal universe than the objects studied by INTEGRAL. This
possibility is supported by the unified model for AGN
as described by Treister & Urry (2005). They predict
a strong correlation of the fraction of broad line AGN
with luminosity, and expect up to a factor of 10 more ab-
sorbed than un-absorbed AGN at very low luminosities
(LX ≃ 1042 erg s−1). This trend can be seen also in the
INTEGRAL AGN sample. When considering only the 14
objects with L20−100 keV < 1043 erg s−1 the ratio of ob-
scured to unobscured objects increases to X = 2.5.
Studying the population of sources at lower limiting
fluxes should also reveal further highly absorbed sources
at high redshifts, because these have been missed so far
and with increasing redshift an increase of absorbed AGN
fraction is expected (La Franca et al. 2005).
The energy range in the 15 -- 200 keV is now also acces-
sible through the BAT instrument aboard Swift (Gehrels
et al. 2005). A study by Markwardt et al. (2005) used
data from the first three months of the Swift mission for
studying the extragalactic sky and reached a flux limit of
f(14−195 keV) ≃ 10−11 erg cm−2 s−1. The source popu-
lation is similar to the INTEGRAL one, with an average
redshift of ¯z = 0.012, and a ratio of X = 2 between
obscured and unobscured AGN, fully consistent with IN-
TEGRAL. Although the Swift/BAT survey covers differ-
ent areas of the sky, the energy band is similar to the IN-
TEGRAL/ISGRI range and the type of AGN detectable
should be the same. Within their sample of 44 AGN they
detect 5 Compton-thick AGN, the same ratio as in the
INTEGRAL sample. The only difference appears in the
relation between luminosity and absorption. While in the
INTEGRAL sample no correlation is detectable (Fig. 5),
Markwardt et al. find in their sample evidence for an anti-
correlation of luminosity and absorption.
4. CONCLUSIONS
The INTEGRAL AGN sample opens the window to the
hard X-ray sky above 20 keV for population studies. With
the 42 extragalactic objects discussed here, a fraction of
about 60% shows absorption above NH = 1022 cm−2,
but only four objects are actually Compton-thick (NH >
1024 cm−2). This shows that the source population above
20 keV, at least at the high flux end, is very similar to the
one observed in the 2 − 10 keV energy region. The re-
sults are consistent with observations by the Swift/BAT
instrument, although we cannot confirm a correlation of
X-ray luminosity with instrinsic absorption. Further in-
vestigations are necessary and with the ongoing INTE-
GRAL and Swift mission it will be revealed in the near
future, if there is a significant Compton-thick AGN pop-
5
ulation to explain the peak in the extragalactic X-ray
background around 30 keV. These objects would have
to have lower fluxes than the objects studied here (i.e.
fX < 10−11 erg cm−2 s−1) and might therefore be low-
redshift, low-luminosity (LX <∼ 1042 erg s−1) Seyfert
galaxies, or higher redshift (z ≫ 0.05) objects.
REFERENCES
Beckmann V., Shrader C. R., Gehrels N., et al. 2005, ApJ
in press, astro-ph/0508327
Beckmann V., Gehrels N., Shrader C. R., & Soldi S.
2006, ApJ accepted, astro-ph/0510530
Comastri A., Setti G., Zamorani G., Hasinger G. 1995,
A&A, 296, 1
Gehrels N., Chincarini G., Giommi P., et al. 2004, ApJ,
611, 1005
George I. M., & Fabian A. C. 1991, MNRAS, 249, 352
Gilli R., Salvati M., Hasinger G., 2001, A&A, 366, 407
Gondek D., Zdziarski A. A., Johnson W. N., et al. 1996,
MNRAS, 282, 646
La Franca F., Fiore F., Comastri A., et al. 2005, ApJ in
press, astro-ph/0509081
Maiolino R., Comastri A., Gilli R., et al. 2003, MNRAS,
344, L59
Markwardt C. B., Tueller J., Skinner G. K., et al. 2005,
ApJL in press, astro-ph/0509860
Pian E., Foschini L., Beckmann V., et al. 2005, A&A,
429, 427
Rothschild R. E., Wilms J., Tomsick J., et al. 2006, ApJ
submitted
Setti G., & Woltjer L. 1989, A&A, 224, L21
Soldi S., Beckmann V., Bassani L., et al. 2005, A&A
accepted, astro-ph/0509123
Sulentic J. W., Marziani P., Zwitter T., et al. 2000, ApJ,
545, L15
Treister E., Urry C. M.,
astro-ph/0505300
2005, ApJ
in press,
Ueda Y., Akiyama M., Ohta K., Miyaji T. 2003, ApJ, 598,
886
Worsley M. A., Fabian A. C., Bauer F. E., et al. 2005,
MNRAS, 357, 1281
Zdziarski A. A., Johnson W. N., Done C., et al. 1995,
ApJ, 438, L63
Table 1. The INTEGRAL AGN sample
Name
z
type
f(20−40 keV)
log L(20−100 keV)
NH
[10−11 erg cm−2 s−1]
[erg s−1]
[1022 cm−2]
6
NGC 788
NGC 1068
NGC 1275
3C 111
MCG +8 -- 11 -- 11
MRK 3
MRK 6
NGC 4051
NGC 4151
NGC 4253
NGC 4388
NGC 4395
NGC 4507
NGC 4593
Coma Cluster
NGC 4945
ESO 323 -- G077
NGC 5033
Cen A
MCG -- 06 -- 30 -- 015
4U 1344 -- 60
IC 4329A
Circinus gal.
NGC 5506
PG 1416 -- 129
IC 4518
NGC 6221
NGC 6300
GRS 1734 -- 292
IGR J18027 -- 1455
ESO 103 -- G35
1H 1934 -- 063
NGC 6814
Cygnus A
MRK 509
IGR J21247+5058
MR 2251 -- 178
MCG -- 02 -- 58 -- 022
Sy 1/2
Sy 2
Sy 2
Sy 1
Sy 1.5
Sy 2
Sy 1.5
Sy 1.5
Sy 1.5
Sy 1.5
Sy 2
Sy 1.8
Sy 2
Sy 1
0.0136
0.0038
0.0176
0.0485
0.0205
0.0135
0.0188
0.0023
0.0033
0.0129
0.0084
0.0011
0.0118
0.0090
0.0231 GClstr.
0.0019
0.0150
0.0029
0.0018
0.0077
0.043
0.0161
0.0014
0.0062
0.1293
0.0157
0.0050
0.0037
0.0214
0.0350
0.0133
0.0106
0.0052
0.0561
0.0344
0.020
0.0634
0.0469
Sy 2
Sy 2
Sy 1.9
Sy 2
Sy 1
?
Sy 1.2
Sy 2
Sy 1.9
Sy 1
Sy 2
Sy 1/2
Sy 2
Sy 1
Sy 1
Sy 2
Sy 1
Sy 1.5
Sy 2
Sy 1
radio gal.?
Sy 1
Sy 1.5
S5 0716+714
S5 0836+710
3C 273
3C 279
PKS 1830 -- 211
BL Lac
0.3a
2.172
FSRQ
0.1583 Blazar
0.5362 Blazar
2.507
Blazar
a tentative redshift
2.98 ± 0.24
0.93 ± 0.27
1.89 ± 0.21
6.27 ± 0.57
6.07 ± 0.97
3.65 ± 0.39
2.01 ± 0.20
1.80 ± 0.20
26.13 ± 0.16
0.93 ± 0.22
9.54 ± 0.25
0.56 ± 0.22
6.46 ± 0.36
3.31 ± 0.16
1.09 ± 0.12
9.85 ± 0.23
1.20 ± 0.19
1.06 ± 0.24
32.28 ± 0.17
2.98 ± 0.19
2.83 ± 0.15
8.19 ± 0.17
10.73 ± 0.18
4.21 ± 0.33
5.43 ± 0.64
0.49 ± 0.32
1.32 ± 0.20
3.91 ± 0.37
4.03 ± 0.09
2.03 ± 0.16
2.97 ± 0.66
0.48 ± 0.25
2.92 ± 0.23
3.24 ± 0.14
4.66 ± 0.47
4.15 ± 0.27
1.20 ± 0.17
1.20 ± 0.28
0.14 ± 0.11
1.73 ± 0.28
5.50 ± 0.15
0.82 ± 0.24
2.07 ± 0.14
43.52
41.92
43.47
44.90
44.06
43.53
43.57
41.60
43.15
42.93
43.55
40.64
43.68
43.08
43.40
42.30
43.21
41.55
42.75
42.93
44.36
44.04
41.97
42.83
45.78
42.92
42.39
42.36
43.88
44.03
43.51
42.51
42.52
44.71
44.42
44.00
44.40
44.18
45.21a
47.87
45.92
46.37
48.09
< 0.02
> 150
3.75
0.634
< 0.02
110
10
< 0.01
4.9
0.8
27
0.15
29
0.02
< 0.01
400
55
2.9
12.5
17.7
2.19
0.42
360
3.4
0.09
?
1
22
3.7
19.0
13 -- 16
?
< 0.05
20
< 0.01
?
0.02 -- 0.19
< 0.01 − 0.08
< 0.01
0.11
0.5
0.02 − 0.13
< 0.01 − 0.7
|
astro-ph/0206483 | 2 | 0206 | 2002-10-04T18:15:36 | How Sensitive Are Weak Lensing Statistics to Dark Energy Content? | [
"astro-ph"
] | Future weak lensing surveys will directly probe the clustering of dark matter, in addition to providing a test for various cosmological models. Recent studies have provided us with the tools which can be used to construct the complete probability distribution function for convergence fields. It is also possible to construct the bias associated with the hot-spots in convergence maps. These techniques can be used in both the quasi-linear and the highly nonlinear regimes using various well developed numerical methods. We use these results here to study the weak lensing statistics of cosmological models with dark energy. We study how well various classes of dark energy models can be distinguished from models with a cosmological constant. We find that the ratio of the square root of the variance of convergence is complementary to the convergence skewness $S_3$ in probing dark energy equation of state; it can be used to predict the expected difference in weak lensing statistics between various dark energy models, and for choosing optimized smoothing angles to constrain a given class of dark energy models. Our results should be useful for probing dark energy using future weak lensing data with high statistics from galaxy weak lensing surveys and supernova pencil beam surveys. | astro-ph | astro-ph |
How Sensitive Are Weak Lensing Statistics
to Dark Energy Content?
Dipak Munshi1,2, and Yun Wang3
1Institute of Astronomy, Madingley Road, Cambridge, CB3 OHA, United Kingdom
2Astrophysics Group, Cavendish Laboratory, Madingley Road, Cambridge, CB3 OHE,
3Department of Physics & Astronomy,University of Oklahoma, Norman, OK 73019 USA.
[email protected], [email protected]
United Kingdom
ABSTRACT
Future weak lensing surveys will directly probe the clustering of dark matter,
in addition to providing a test for various cosmological models. Recent studies
have provided us with the tools which can be used to construct the complete
probability distribution function for convergence fields.
It is also possible to
construct the bias associated with the hot-spots in convergence maps. These
techniques can be used in both the quasi-linear and the highly nonlinear regimes
using various well developed numerical methods. We use these results here to
study the weak lensing statistics of cosmological models with dark energy. We
study how well various classes of dark energy models can be distinguished from
models with a cosmological constant. We find that the ratio of the square root
of the variance of convergence is complementary to the convergence skewness S3
in probing dark energy equation of state; it can be used to predict the expected
difference in weak lensing statistics between various dark energy models, and for
choosing optimized smoothing angles to constrain a given class of dark energy
models. Our results should be useful for probing dark energy using future weak
lensing data with high statistics from galaxy weak lensing surveys and supernova
pencil beam surveys.
Subject headings: Cosmology: theory -- weak lensing -- Methods: analytical --
Methods: statistical -- Methods: numerical
1.
Introduction
Recent cosmological observations favor an accelerating universe (Garnavich et al. 1998a;
Riess et al. 1998; Perlmutter et al. 1999). This implies the existence of energy of unknown
-- 2 --
nature (dark energy), which has negative pressure. Current data are consistent with dark en-
ergy being a non-zero cosmological constant (see for example, Wang & Garnavich 2001; Bean
& Melchiorri 2002). Many other alternative dark energy candidates have been considered,
and are consistent with data as well. For example, quintessence, k-essence, spintessence, etc.
(Peebles & Ratra 1988; Frieman et al. 1995; Caldwell, Dave, & Steinhardt 1998; Garnavich
et al. 1998b; White 1998; Efstathiou 1999; Steinhardt, Wang, & Zlatev 1999; Podariu &
Ratra 2000; Sahni & Wang 2000; Sahni & Starobinsky 2000; Saini et al. 2000; Waga &
Frieman 2000; Huterer & Turner 2001; Ng & Wiltshire 2001; Podariu, Nugent, & Ratra
2001; Weller & Albrecht 2001)
Different dark energy models can be conveniently classified according to the equation of
state of the dark energy component, wX. For example, for quintessence models, dwX/dz >
0, while for k-essence models, dwX/dz < 0. There are many complimentary probes of
dark energy. These include, the distance-redshift relations of cosmological standard candles,
Cosmic Microwave Background Anisotropy, volume-redshift test using galaxy counts, the
evolution of galaxy clustering, weak lensing, etc. These different methods to probe dark
energy are complimentary to each other, and can provide important consistency checks, due
to the different sources of systematics in each method (for example, see Kujat et al. 2002).
Weak lensing surveys (Bacon, Refregier & Ellis 2000; Van Waerbeke et al.
2000;
Wittman et al. 2000; Maoli et al. 2001; Van Waerbeke et al. 2001; Wilson, Kaiser, &
Luppino 2001; Bacon et al. 2002; Hoekstra et al. 2002; Refregier, Rhodes & Groth 2002),
currently underway and more proposed in the near future, are well suited to studying the
dark energy equation of state. Weak lensing directly probes the gravitational clustering and
the background cosmology. Many recent studies, both theoretical and numerical, have ana-
lyzed these possibilities. Observational teams have already reported first detections of cosmic
shear. On theoretical front progress has been made in modeling the statistics using both
perturbative calculations which are valid for large smoothing angles and also using the well
motivated hierarchical ansatz which is valid for small smoothing angles. Numerical studies
carried out so far uses ray shooting experiments and are quite useful in testing analytical
calculations.
In an earlier study on probing quintessence using weak lensing, Hui (1999) concluded
that the large scale convergence skewness can directly provide a constraint for wX, the equa-
tion of state of dark energy. Similarly, Huterer studied the use of weak lensing convergence
power spectrum and three-point statistics to constrain dark energy models. Both of these
papers are useful in utilizing large weak lensing surveys of galaxies to constrain dark energy.
In this paper we construct the complete probability distribution function of convergence
to study the effects of dark energy. Our results apply to the weak lensing of both galaxies
-- 3 --
(on large angular scales) and type Ia supernovae (on small angular scales). We compare two
classes of dark energy models (one with effective constant equation of state wX, the other
with time-varying wX) against that of a Λ dominated universe.
The technique we use in this paper has been tested in detail using N-body calculations.
Analytical results are obtained for large smoothing angles using the perturbative calculations,
and for small smoothing angles using the hierarchical ansatz. We focus on both the one-
point probability distribution function and the bias associated with convergence maps in
the quasi-linear and the highly nonlinear regimes. Our studies are quite complementary
to the studies done using a Fisher matrix analysis for the recovery of power spectra from
observations. While such studies are well suited for recovering the nonlinear matter power
spectra; the study of the probability distribution function and the bias will give us a direct
probe of non-Gaussianity developed through gravitational clustering.
The paper is organized as follows.
In Section 2 we give the definition of some basic
equations for reference, and identify the dark energy models studied in this paper. Section
3 discusses the weak lensing statistics of dark energy models. Section 4 contains discussions
and a summary.
2. Notation
The weak lensing convergence, κ, maps the distribution of projected density fields, and its
statistics is directly related to that of the underlying matter distribution. We write
κ(θ0) = Z χs
0
dχω(χ)δ(r(χ)θ0, χ),
(1)
where r(χ) is the angular diameter distance and ω(χ) is the weight function associated with
the source distribution. The observer is placed at χ = 0, and the sources (which for simplicity
we assume are all at the same redshift) are placed at χs. χ is given by
χ(z) = cH−1
0 Z z
0
dz′[Ωm(1 + z′)3 + Ωk(1 + z′)2 + ΩX f (z)]−1/2
(2)
where the Ω's denote the fraction of critical density in various components. ΩX denotes
the dark energy component, and Ωk = 1 − Ωm − ΩX , The function f (z) parametrizes the
time-dependence of the dark energy density, and f (z = 0) = 1. For dark energy with
constant equation of state, wX = pX/ρX = constant, f (z) = (1 + z)3(1+wX ). The limiting
case with wX = −1 [i.e., f (z) = 1] corresponds to a cosmological constant. Note that to
-- 4 --
obtain accelerated expansion of the universe, we need ρ + 3p < 0, which implies wX < −1/3
for a power law dark energy density f (z). In general, the dark energy equation of state,
wX(z), can be written in terms of the dimensionless dark energy density, f (z), wX(z) =
1
3(1 + z) f ′(z)/f (z) − 1, for dark energy density or equation of state with arbitrary time
dependence. (Wang & Garnavich 2001)
We study two classes of dark energy models. The first class contains dark energy models
with effective constant equation of state, wX = −1/3, −2/3, −1 (ΛCDM), and −1.9. If dark
energy arises from classical fields, it must satisfy the weak energy condition, which requires
that wX > −1. The weak energy condition violating toy model, wX = −1.9, is motivated
by quantum gravity models of inflation in which quantum effects lead to the violation of
the weak energy condition (Onemli & Woodard 2002). The second class contains two dark
energy toy models with linear time-varying equations of state, wX = wq(z) = −1 + z and
−1 + 2z/3. These are motivated by quintessence models which have wX that effectively
increases with z. All the models have Ωm = 0.3, ΩX = 0.7, h = 0.7, nS = 1 (power law
index of the primordial power spectrum), and σ8 = 0.8. We normalize the nonlinear power
spectrum to σ8. Table 1 lists these dark energy models.
Table 1
Dark Energy Models
Model
ΛCDM
constant wX −1/3,
wq(z)
−1
wX (z)
−2/3,
−1 + 2z/3, −1 + z
−1.9
The two classes of dark energy models are compared with the fiducial ΛCDM model to
quantify the variation in various statistics of convergence maps.
3. Statistics of Weak Lensing in Dark Energy Models
To compute the statistics of weak lensing convergence field, we need to relate it to
the statistics of three dimensional density field of underlying matter distribution. In recent
studies such analysis has been done extensively. We use such a formalism to explore the
weak lensing statistics for the dark energy cosmologies. For large smoothing angles we use
the perturbative calculations and for small smoothing angles we use the well motivated
hierarchical ansatz to compute the relevant quantities.
-- 5 --
3.1. Evolution of the Matter Power Spectrum
We compute the matter power spectrum using the scaling ansatz of Hamilton et al.
(1991), which was later extended by various authors [see e.g. Peacock & Dodds (1996)].
This ansatz essentially consists of postulating that 4πk3P (k) = f [4πk3
l Pl(kl)], where P (k)
is the nonlinear power spectrum and Pl is the linear power spectrum, and the function f in
general will depend on the initial power spectrum. The linear power spectrum is evaluated
at a different wave number, kl = (1 + 4πk3P (k))−1/3k, hence the mapping is non-local in
nature. The cosmological model enters through the linear growth function g(z), so that
Pl(k, z) = [g(z)/(1 + z)]2 Pl(k, z = 0). The linear growth function is given by:
g(z) =
δ(z, Ωm, ΩΛ)
δ(z, Ωm = 1)
=
5
2
Ωm (1 + z) E(z) Z ∞
z
dz′
(1 + z′)
[E(z′)]3 ,
Where
E(z) ≡ pΩm(1 + z)3 + Ωk(1 + z)2 + ΩX f (z)
(3)
(4)
We compute the linear growth function by direct integration [for more on power spectrum
evolution in quintessence models see (Benabed & Bernardeau 2001)].
Our method enforces stable clustering in the nonlinear regime and assumes the hier-
archical ansatz (which is tested by numerous numerical simulations), therefore we are able
to predict the higher order correlation functions (Davis & Peebles 1977, Groth & Peebles
1977, Fry & Peebles 1978) and (their Fourier transforms or) the multi-spectrum correctly.
Combining these with the powerful technique of the generating function we can construct the
complete probability distribution function and the bias associated with convergence maps.
3.2. Convergence Probability Distribution Function
Perturbative calculations depend on the expansion of the convergence field κ(θ0) for
smoothing angle θ0 in terms of perturbative expansion of the density field δ. Such an analysis
can be performed in an order by order manner,
κ(1)(θ0) + κ(2)(θ0) + . . . = Z χs
0
dχω(χ) (cid:2)δ(1)(r(χ)θ0) + δ(1)(r(χ)θ0) + . . .(cid:3)
(5)
-- 6 --
where δ(i) and κ(i) correspond to the i-th order perturbative expansion, i = 1 being the
linear order.
In the perturbative regime at tree level (Fry 1984; Bernardeau 1992, 1994;
Bernardeau & Schaeffer 1992), it is possible to introduce vertex generating function G(τ )
which will encode the tree level contribution from all orders. The smoothing using a top-
hat filter function can be incorporated in the generating function formalism and then the
generating function can be written in terms of the generating function of the unsmoothed
case. All statistical quantities including probability distribution functions can be constructed
once we have solved for the tree-level generating functions (Bernardeau 1992, 1994). We write
G(τ ) = (cid:18)1 −
τ
κa(cid:19)−κa
; GP T
s
(τ ) = GP T (cid:20)τ
σ(R0(1 + GP T (τ ))1/2)
σ(R0)
(cid:21)
(6)
The parameter κa can be determined from the dynamical equation governing the evolution
of perturbations in the quasi-linear regime, and it is given by κa =
. The variance at a
length scale σ(R0) = R−(n+3)/2
, where a local power law spectrum index n is used to evaluate
the generating function. The generating function is now used to compute the probability
distribution function at a particular smoothing scale.
√13−1
0
2
In the highly nonlinear regime (Balian & Schaeffer 1989, Davis & Peebles 1977, Groth
& Peebels 1977, Szapudi & Szalay 1993, Scoccimarro & Frieman 1999, Munshi et al. 1999c),
the perturbative series starts to diverge and the usual perturbative calculations are replaced
by the hierarchical ansatz for higher order correlation functions, which can be built from
two-point correlation functions. The amplitude of various contributions can be constructed
from the knowledge of the generating function.
It was found from analytical reasoning
and numerical experimentation that the generating function in the highly nonlinear regime
retains exactly the same form as in the quasi-linear regime; however, the value of κa is
changed (Beranrdeau 1992) -- it is now treated as a free parameter (Colombi et al. 1995,
It is customary to use a different parameter ω that is easy to
Munshi et al. 1999a,b).
evaluate from numerical simulations, κa = 2ω
It was found that ω = .3 reproduces
various statistics in the nonlinear regime quite well (Colombi et al. 1997, Colombi, Bouchet,
Schaeffer 1995, Munshi et al. 1999d). To compute the probability distribution function, one
has to compute the void probability function φ(y), which acts as a generating function for
normalized cumulants or SN parameters and can be expressed in terms of the function G(τ )
as (Balian & Schaeffer 1989):
(1−ω) .
1
2
yτ
d
dτ
G(τ )
φ(y) = yG(τ ) −
τ = −y
d
dτ
G(τ ).
(7)
-- 7 --
Finally the probability distribution function can now be written as (Balian & Schaeffer 1989):
P (δ) = Z i∞
−i∞
dy
2πi
exp(cid:20) (1 + δ)y − φ(y)
¯ξ2
(cid:21)
(8)
In recent studies it was found that in both the quasi-linear and the highly nonlinear
regimes, it is possible to introduce a reduced convergence field (Munshi & Jain 2000, Munshi
2002), η = κ−κmin
, which to a very good approximation follows the same statistics as 1+δ. In
−κmin
the quasi-linear regime it follows the smoothed projected density, and in the highly nonlinear
regime it simply follows the 3D statistics of the density field.1
The variance of η is given by (Valageas 2000a; Valageas 2000b)
ξη = Z χs
0
dχ (cid:18) w
Fs(cid:19)2
Iµ(χ),
with
w(χ, χs) =
H 2
0
c2
D(χ) D(χs − χ)
D(χs)
(1 + z),
D(χ) =
Fs = Z χs
0
dχ w(χ, χs),
Iµ(z) = πZ ∞
0
dk
k
∆2(k, z)
k
sinn(cid:16)pΩk χ(cid:17)
cH−1
0
pΩk
W 2(Dkθ0),
(9)
(10)
where "sinn" is defined as sinh if Ωk > 0, and sin if Ωk < 0.
If Ωk = 0, the sinn and
Ωk's disappear. ∆2(k, z) = 4πk3P (k, z), k is the wavenumber, P (k, z) is the matter power
spectrum, The window function W (Dkθ0) = 2J1(Dkθ0)/(Dkθ0) for smoothing angle θ0. Here
J1 is the Bessel function of order 1. Note that the clustering of the dark energy field would
lead to an increase in the transfer function on very large scales. Huterer (2002) has shown
that the clustering of the dark energy field can be neglected on the scales relevant to weak
lensing surveys.
Fig.1 shows −κmin and pξη, and Figs. 2 & 3 show the pdf for the two classes of dark
energy models listed in Table 1.
1It was recently shown by Munshi (2002) that analytical results obtained by direct perturbative calcula-
tions can also be obtained by using a functional fit obtained from assuming a log-normal evolution of local
correlated density field. This method was also found not only to work for one point probability distribution
function but also for bias associated with convergence maps.
-- 8 --
3.3. Bias Associated with Convergence Maps
Assuming a correlation function hierarchy guarantees that we have a two-point proba-
bility distribution function P (κ1, κ2) which can be factorized as follows (Munshi 2001),
P (κ1, κ2)dκ1dκ2 = P (κ1)P (κ2) [1 + b(κ1)ξκ
12b(κ2)] dκ1dκ2.
(11)
The function ξκ
12 is the two point correlation function corresponding to convergence maps.
The function b(κ) is the bias associated with the convergence maps, and can be shown to be
related to the bias associated with overdense regions, b(1 + δ) (Munshi 2001). The perturba-
tive calculations also produce similar results, although the nature of the bias function b(κ)
changes from one regime to another. As was the case for one point normalized moments,
whose generating function was related to the one point probability distribution function, the
generating functions for two-point collapsed higher order correlation functions, also known
as cumulant correlators, are also related to the bias function in a very similar manner. We
write
P (δ)b(δ) = Z i∞
−i∞
dy
2πi
τ exp(cid:20) (1 + δ)y − φ(y)
¯ξ2
(cid:21) ,
(12)
where τ is a generating function for the cumulant correlators. However, it turns out that
the differential bias, as we have written down above, is difficult to estimate from numerical
data. Therefore we work with the cumulative bias, which is the bias associated with points
where convergence maps cross a particular threshold (Munshi 2001). Previous studies against
numerical simulations showed that the bias function describes the numerical results quite
accurately. It was shown in earlier studies that b(κ) = b(1+δ)
κmin
.
Figures 4 & 5 show the cumulative bias for the two classes of dark energy models
indicated in Figure 1. Clearly, the cumulative bias of convergence is complementary to the
convergence pdf in probing the non-Gaussianity of gravitational clustering and constraining
dark energy models.
3.4. A New Indicator for Deviations from the ΛCDM model
We find that the deviations of dark energy models from the fiducial ΛCDM model can
be quantified with a single parameter
-- 9 --
1 + ǫ ≡ p ¯ξκ(XCDM)
p ¯ξκ(ΛCDM)
=
κmin(XCDM)p ¯ξη(XCDM)
κmin(ΛCDM)p ¯ξη(ΛCDM)
,
(13)
where XCDM represents an arbitrary dark energy model. Figure 6 shows the indicator 1 + ǫ
for the two classes of dark energy models studied in this paper, for smoothing angle θ0 = 1′
and 15′.
Comparison of Fig.6 and Figs.2-3 shows that the more the pdf of the dark energy model
differs from that of the fiducial ΛCDM model, the more the indicator 1 + ǫ deviates from
one. This indicates that the pdf is primarily determined by its variance, which is consistent
with the finding of the existence of a universal probability distribution function (in terms of
the scaled convergence η) for weak lensing amplification by Wang, Holz, & Munshi (2002).
It is useful to compare our new indicator with the convergence skewness S3 (Hui 1999)
for the same models. Figure 7 shows S3 for the same models as in Figure 6, with the
same line and arrow types. We have computed the convergence skewness S3 using Hyper-
Extended Perturbation theory in the nonlinear regime (small smoothing angular scales),
and perturbative results are adopted for the quasi-linear regime (larger smoothing angular
scales). While S3, as an indicator, mainly encodes the information about non-Gaussianity,
the indicator we have proposed is directly related to the variance and is more sensitive to
the projected density power spectrum.
Our new weak lensing pdf shape indicator, 1 + ǫ, is complementary to S3 in constraining
the dark energy equation of state. The indicator, 1 + ǫ, is sensitive to smoothing angle θ0,
while S3 is not very sensitive to θ0 at small angular scales. Feasible future supernova surveys
can yield a large number of type Ia supernovae out to redshift z = 1 and beyond (Wang
2000, SNAP2). It may be possible to directly measure the weak lensing pdf with sufficiently
high statistics (Metcalf & Silk 1999; Seljak & Holz 1999); this would allow us to utilize the
pdf with different smoothing angles to probe different ranges of constant wX models, and
the variation of wX with z.
4. Discussions and Summary
We have analyzed weak lensing statistics for two classes of dark energy cosmological
models. One class of dark energy models have effective constant equation of state wX , while
the other have linear time-varying wX(z) inspired by quintessence models. The weak lensing
2See http://snap.lbl.gov.
-- 10 --
statistics of these dark energy models are compared with that of a ΛCDM model.
It has been shown that in directly using the distance-redshift relations of type Ia su-
pernovae to probe dark energy, it is optimal to measure the dark energy density, ρX (z) =
ρX (0) f (z), instead of the dark energy equation of state, wX(z). (Wang & Garnavich 2001;
Wang & Lovelace 2001; Tegmark 2001) In this paper, for convenience and illustration, we
have used wX to classify various models.
Note that we have considered a dark energy toy model which violates the weak energy
condition, as similar models could arise from quantum effects in quantum gravity models
of inflation (Onemli & Woodard 2002). Also, we have only considered dark energy models
with linear time-varying wX (z), although dark energy models with much more complicated
time dependence in wX (z) have been proposed (see for example, Bassett et al. 2002). This
is because it is extremely difficult to extract the time dependence of wX(z), even if it were
a simple linear function of the redshift z, from observational data (for example, see Maor et
al. 2001; Barger & Marfatia 2001; Wang & Garnavich 2001; Kujat et al. 2002; Maor et al.
2002).
We have studied the statistics of the cosmic convergence field via various diagnostics
including the one point probability distribution functions and the bias associated with conver-
gence "hot spots". The analysis was done for both the quasi-linear scales where perturbative
calculations are valid and also for very small angular scales where hierarchical ansatz is gen-
erally used to quantify the statistical distributions. Following earlier studies we introduce a
quantity κmin which can help us to write the observed convergence field in terms of a reduced
convergence field which in turn represents directly the statistics of density distribution. For
large smoothing angles, perturbation theory predicts this quantity η to trace the projected
density field, and on small angular scales it traces the nonlinear density field in three di-
mensions. The lower order moments corresponding to various cosmologies have already been
investigated in detail and our studies complement these results. Also we have used top-hat
window functions, but Bernardeau & Valageas (2000) have shown how to generalize similar
calculations in terms of aperture mass using compensated filters.
We have identified a new weak lensing pdf shape indicator, 1 + ǫ ∝ p ¯ξκ = κminp ¯ξη,
which can be used to predict the expected difference in weak lensing statistics between various
dark energy models, and for choosing optimized smoothing angles to constrain a given class
of dark energy models. For example, small smoothing angles are favored for constraining
dark energy models with wX < −1.
Our proposed 1 + ǫ statistics is related to the volume average of two-point statistics
of ¯ξκ. Note that while ¯ξη only depends on the smoothing angle and the underlying mass
-- 11 --
distribution, κmin encodes the dependence on cosmological parameters. Therefore, our new
statistical indicator, 1 + ǫ ∝ p ¯ξκ = κminp ¯ξη, is of interest as far as we are interested in
differentiating various cosmological models. We found that both P (κ) and b(κ) depends
on kmin and ξη, however, it is difficult to infer the difference in these statistics in various
cosmologies. The 1 + ǫ statistics we have devised, however, is quite interesting in the sense
that it can very easily be used to check how much various dark energy models differ in weak
lensing. It can be used to complement and supplement various other statistics such as S3.
We have computed the convergence skewness S3 for the dark energy models considered
in this paper. We find that the new weak lensing pdf shape indicator, 1 + ǫ, is indeed
complementary to S3 in probing dark energy equation of state.
We note that getting maps of convergence is difficult compared to direct evaluation of
non-Gaussian statistics from shear maps (see e.g. Schneider & Lombardi 2002; Zaldariaga
& Scoccimarro 2002). On the other hand, the construction of convergence statistics can be
directly modeled at arbitrary level, whereas for shear field the computation of statistics is
done in a order by order manner so far. So an independent analysis of convergence maps
constructed from shear maps should be useful in constraining various errors which might get
introduced during various stages of data reduction. The question of error bars in weak lensing
measurements has been dealt with in great detail in Munshi & Coles (astro-ph/0003481, to
appear in MNRAS) for various window functions and are independent of the cosmological
model assumed. Our convergence statistics can be a powerful diagnostic and complementary
tool to the shear map statistic.
With high statistics data from future weak lensing surveys of galaxies and supernova
pencil beam surveys (Wang 2000; SNAP), weak lensing can be a useful tool in differentiating
different dark energy models.
DM was supported by PPARC grant RG28936, and YW was supported in part by NSF
CAREER grant AST-0094335. DM would like to thank Alexandre Refregerier for many
useful discussion, and Francis Bernardeau for making a copy of his code to compute the pdf
and bias available to us.
Bacon, D.J.; Refregier, A.R.; Ellis, R.S. 2000, MNRAS, 318, 625
REFERENCES
Bacon, D.J.; Massey, R.; Refregier, A.R.; Ellis, R.S. 2002, astro-ph/0203134, submitted to
MNRAS
-- 12 --
Benabed,K., Bernardeau, F., Phys.Rev. D64 (2001) 083501
Barger, V., & Marfatia, D. 2001, Phys. Lett. B, 498, 67
Balian R., Schaeffer R., 1989, A& A, 220, 1
Bartelmann, M., Schneider, P., 1991, A& A, 248, 353
Bassett, B. A., Kunz, M., Silk, J., & Ungarelli, C. 2002, astro-ph/0203383
Bean, R., & Melchiorri, A., Phys.Rev. D65 (2002) 041302
Bernardeau, F., 1992, ApJ, 392, 1
Bernardeau, F., 1994, A& A, 291, 697
Bernardeau, F., 1998, A& A, 338, 375
Bernardeau, F., Schaeffer, R., 1992, A& A, 255, 1
Bernardeau, F., van Waerbeke, L., Mellier, Y., 1997, A& A, 322, 1
Bernardeau, F.; Valageas, P. 2000, A & A, 364, 1
Blandford, R.D., Saust, A.B., Brainerd, T.G., Villumsen, J.V., 1991, MNRAS, 251, 600
Boschan, P., Szapudi, I., Szalay, A.S. 1994, ApJS, 93, 65
Caldwell, R.R., Dave, R., Steinhardt, P.J. 1998, Phys. Rev. Lett. 80, 1582
Colombi, S., Bouchet, F.R.,Schaeffer, R., 1995, ApJS, 96, 401
Colombi, S., Bernardeau, F., Bouchet, F.R., Hernquist, L., 1997, MNRAS,287, 241
Davis, M., Peebles, P.J.E., 1977, ApJS, 34, 425
Efstathiou, G. 1999, MNRAS, 310, 842
Frieman, J.A.; Hill, C.T.; Stebbins, A.; & Waga, I. 1995, Phys. Rev. Lett. 75, 2077
Fry, J.N., 1984, ApJ, 279, 499
Fry, J.N., Peebles, P.J.E., 1978, ApJ, 221, 19
Garnavich, P.M. et al. 1998a, ApJ, 493, L53
Garnavich, P.M. et al. 1998b, ApJ, 509, 74
-- 13 --
Groth, E., Peebles, P.J.E., 1977, ApJ, 217, 385
Hamilton, A.J.S., Kumar, P., Lu, E., Matthews, A., 1991, ApJ, 374, L1
Hoekstra, H., et al. 2002, ApJ, 572, 55
Hui, L., 1999, ApJ, 519, L9
Huterer, D., & Turner, M. S. 2001, Phys. Rev. D64, 123527
Huterer, D. 2002, Phys. Rev. D65, 063001
Jain, B., Mo H.J., White S.D.M., 1995, MNRAS, 276, L25
Jain, B., Seljak, U., 1997, ApJ, 484, 560
Jain, B., van Waerbeke L., 2000, ApJ, 530, L1
Kaiser, N., 1992, ApJ, 388, 272
Kaiser, N., 1998, ApJ, 498, 26
Kujat, J., Linn, A.M., Scherrer, R.J., Weinberg, D.H. 2002, ApJ, 572, 1-14,2001
Maoli, R., et al. 2001, A & A, 368, 766
Maor, I., Brustein, R., & Steinhardt, P.J. 2001, Phys. Rev. Lett., 86, 6; Erratum-ibid. 87
(2001) 049901
Maor, I., Brustein, R., McMahon, J., & Steinhardt, P.J. 2002, Phys.Rev. D65 (2002) 123003
Metcalf, R. B., & Silk, J. 1999, ApJ, 519, L1
Munshi, D., Coles, P., Melott, A.L., 1999a, MNRAS, 307, 387
Munshi, D., Coles, P., Melott, A.L., 1999b, MNRAS, 310, 892
Munshi, D., Melott, A.L., Coles, P., 1999c, MNRAS, 311, 149
Munshi, D., Bernardeau, F., Melott, A.L., Schaeffer, R., 1999d, MNRAS, 303, 433
Munshi, D., Coles, P., 2000a, MNRAS, 313, 148
Munshi, D., Coles, P., 2000b, MNRAS, submitted
Munshi, D., Coles, P., 2002, MNRAS in press, astro-ph/0003481
-- 14 --
Munshi, D., Jain, B., 2000, MNRAS, 318, 109
Munshi, D, 2001, MNRAS, 322, 107
Ng, S.C.C., & Wiltshire, D.L. 2001, Phys. Rev. D64, 123519
Onemli, V. K., Woodard, R. P. 2002, gr-qc/0204065.
Peacock, J.A., Dodds, S.J., 1996, MNRAS, 280, L19
Peebles, P. J. E.; & Ratra, B. 1988, ApJ, 325L, 17
Perlmutter, S., et al. 1999, ApJ, 517, 565
Podariu, S., & Ratra, B. 2000, ApJ, 532, 109
Podariu, S., Nugent, P., & Ratra, B. 2001, ApJ, 553, 39
Refregier, A., Rhode, J, & Groth, E., 2002, ApJ, 572, L131
Rhode, J, Refregier, A., Groth, E., 2001, ApJ, 552, L85
Riess, A.G., et al 1998, AJ, 116, 1009
Sahni, V., & Starobinsky, A. 2000, Int. J. Mod. Phys. D, 9, 373
Sahni, V., & Wang, L. 2000, Phys. Rev. D, 62, 103517
Saini, T.; Raychaudhury, S.; Sahni, V.; and Starobinsky, A.A. 2000, Phys. Rev. Lett., 85,
1162
Schneider, P., & Lombardi, M. 2002, astro-ph/0207454, submitted to A & A
Scoccimarro R., & Frieman J.A. 1999, ApJ, 520, 35
Seljak, U., & Holz, D. E. 1999, A&A, 351, L10
Steinhardt, P.J., Wang, L., & Zlatev, I. 1999, Phys. Rev. D59, 123504
Szapudi, I., Szalay, A.S., 1993, ApJ, 408, 43
Tegmark, M. 2001, astro-ph/0101354
Valageas, P., 2000a, A& A, 354, 767
Valageas, P., 2000b, A& A, 356, 771
-- 15 --
Van Waerbeke, L. et al. 2000, A & A, 358, 30
Van Waerbeke, L., et al. 2001, A & A, 374, 757
Waga, I., & Frieman, J.A. 2000, Phys.Rev. D62, 043521
Wang Y., 1999, ApJ, 525, 651
Wang, Y. 2000, ApJ, 531, 676
Wang, Y. & Garnavich, P. 2001, ApJ, 552, 445
Wang, Y. & Lovelace, G. 2001, ApJ, 562, L115
Wang Y., Holz, D.E., & Munshi, D. 2002, ApJ, 572, L15
Weller, J., & Albrecht, A. 2001, astro-ph/0106079
White, M. 1998, ApJ, 506, 495
Wilson, G.; Kaiser, N.; Luppino, G.A. 2001, ApJ, 556, 601
Wittman, D.M, Tyson, J.A., Kirkman, D., Dell'Antonio,I., Bernstein, G., 2000, nature, 405,
143
Zaldarriaga, M., & Scoccimarro, R. 2002, astro-ph/0208075
This preprint was prepared with the AAS LATEX macros v5.0.
-- 16 --
Fig. 1. -- The minimum value of the convergence field, κmin, and variance of the scaled
convergence η = 1 + κ/κmin, ξη, as functions of constant dark energy equation of state,
wX, for source redshift zs = 0.5 and 1.0. The arrows indicate the −κmin and pξη values for
dark energy models with time-varying equation of state, wq(z) = −1 + z (long arrows) and
wq(z) = −1 + 2z/3 (short arrows). Note that κmin does not depend on the smoothing angle
θ0 but it depends on the background dynamics of the universe.
Fig. 2. -- PDF associated with constant wX models. Two different regimes are considered
for computing the PDF. Upper panels θs = 1′, 4′ correspond to the nonlinear calculations
where we have assumed a hierarchical ansatz for the correlation hierarchy. For the lower
panels perturbative results are used to construct the PDF at smoothing angles θs = 15′
and 60′. Various curves correspond to various quintessence models.
In each panel solid
line represents the ΛCDM model, short dashed line represent wX = −2/3, dot-dashed line
represents wX = −1/3, and long dashed line represents wX = −1.9.
Fig. 3. -- PDF associated with time-varying wX models compared to ΛCDM. As in the
previous figure two different regimes are considered for computing the PDF. Upper panels
θs = 1′, 4′ correspond to the nonlinear calculations where we have assumed a hierarchical
ansatz for the correlation hierarchy. For the lower panels perturbative results are used to
construct the PDF at smoothing angles θs = 15′ and 60′. Various curves correspond to
various quintessence models.
In each panel solid line represents the ΛCDM model, short
dashed line represent wX = −1 + z model, and long dashed line represents wX = −1 + 2z/3
model.
Fig. 4. -- Bias associated with constant wX models. Two different regimes are considered
for computing the PDF. Upper panels θs = 1′, 4′ correspond to the nonlinear calculations
where we have assumed a hierarchical ansatz for the correlation hierarchy. For the lower
panels perturbative results are used to construct the PDF at smoothing angles θs = 15′
and 60′. Various curves correspond to various quintessence models.
In each panel solid
line represents the ΛCDM model, short dashed line represent wX = −2/3, dot-dashed line
represents wX = −1/3, and long dashed line represents wX = −1.9.
Fig. 5. -- Bias associated with time-varying wX models compared to ΛCDM. As in previous
figure two different regimes are considered for computing the PDF. Upper panels θs = 1′, 4′
correspond to the nonlinear calculations where we have assumed a hierarchical ansatz for
the correlation hierarchy. For the lower panels perturbative results are used to construct the
PDF at smoothing angles θs = 15′ and 60′. Various curves correspond to various quintessence
models. In each panel solid line represents the ΛCDM model, short dashed line represent
wX = −1 + z model, and long dashed line represents wX = −1 + 2z/3 model.
-- 17 --
Fig. 6. -- Indicator of the deviations of dark energy models from the fiducial ΛCDM model,
1 + ǫ, as function of constant dark energy equation of state wX , for smoothing angle θ0 = 1′
and 15′. The arrows indicate the 1 + ǫ values for dark energy models with time-varying
equation of state, wq(z) = −1 + z (long arrows), and wq(z) = −1 + 2z/3 (short arrows).
Fig. 7. -- The skewness S3, for the same models as in Fig.6. The line and arrow types are
the same.
In the left panel we use Hyper-Extended Perturbation theory to compute the
convergence skewness, whereas in the right panel (where larger smoothing angular scales are
considered) perturbative results are adopted.
-- 18 --
0.6
0.4
0.2
0.1
0.08
0.06
0.04
0.02
0
-2
-1.5
-1
-0.5
0
-2
-1.5
-1
-0.5
Fig. 1. -- The minimum value of the convergence field, κmin, and variance of the scaled
convergence η = 1 + κ/κmin, ξη, as functions of constant dark energy equation of state,
wX, for source redshift zs = 0.5 and 1.0. The arrows indicate the −κmin and pξη values for
dark energy models with time-varying equation of state, wq(z) = −1 + z (long arrows) and
wq(z) = −1 + 2z/3 (short arrows). Note that κmin does not depend on the smoothing angle
θ0 but it depends on the background dynamics of the universe.
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0.001
0.001
0.0001
0.0001
100
100
100
100
10
10
10
10
1
1
1
1
0.1
0.1
0.1
0.1
0.01
0.01
0.01
0.01
0
0
0.08
0.08
0.16
0.16
0.001
0.001
0.001
0.001
-0.02
-0.02
-0.02
-0.02
0
0
0
0
0.02
0.02
0.02
0.02
-- 19 --
100
10
1
0.1
0.01
0.001
0.0001
100
100
100
100
10
10
10
10
1
1
1
1
0.1
0.1
0.1
0.1
0.01
0.01
0.01
0.01
0.001
0.001
0.001
0.001
0
0.04
0.08
-0.01
-0.01
-0.01
-0.01
0
0
0
0
0.01
0.01
0.01
0.01
Fig. 2. -- PDF associated with constant wX models. Two different regimes are considered
for computing the PDF. Upper panels θs = 1′, 4′ correspond to the nonlinear calculations
where we have assumed a hierarchical ansatz for the correlation hierarchy. For the lower
panels perturbative results are used to construct the PDF at smoothing angles θs = 15′
and 60′. Various curves correspond to various quintessence models.
In each panel solid
line represents the ΛCDM model, short dashed line represent wX = −2/3, dot-dashed line
represents wX = −1/3, and long dashed line represents wX = −1.9.
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0.001
0.001
0.0001
0.0001
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0
0
0.08
0.08
0.16
0.16
0.001
0.001
-0.02
-0.02
0
0
0.02
0.02
-- 20 --
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0.001
0.001
0.0001
0.0001
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0.001
0.001
0
0
0.04
0.04
0.08
0.08
-0.01
-0.01
0
0
0.01
0.01
Fig. 3. -- PDF associated with time-varying wX models compared to ΛCDM. As in the
previous figure two different regimes are considered for computing the PDF. Upper panels
θs = 1′, 4′ correspond to the nonlinear calculations where we have assumed a hierarchical
ansatz for the correlation hierarchy. For the lower panels perturbative results are used to
construct the PDF at smoothing angles θs = 15′ and 60′. Various curves correspond to
various quintessence models.
In each panel solid line represents the ΛCDM model, short
dashed line represent wX = −1 + z model, and long dashed line represents wX = −1 + 2z/3
-- 21 --
model.
0
0.08
0.16
1000
100
10
1000
100
0
0.04
0.08
0
0.02
10
-0.01
0
0.01
0.02
100
10
1000
100
10
Fig. 4. -- Bias associated with constant wX models. Two different regimes are considered
for computing the PDF. Upper panels θs = 1′, 4′ correspond to the nonlinear calculations
where we have assumed a hierarchical ansatz for the correlation hierarchy. For the lower
panels perturbative results are used to construct the PDF at smoothing angles θs = 15′
and 60′. Various curves correspond to various quintessence models.
In each panel solid
line represents the ΛCDM model, short dashed line represent wX = −2/3, dot-dashed line
-- 22 --
represents wX = −1/3, and long dashed line represents wX = −1.9.
0
0.08
0.16
1000
100
10
1000
100
0
0.04
0.08
0
0.02
10
-0.01
0
0.01
0.02
100
10
1000
100
10
Fig. 5. -- Bias associated with time-varying wX models compared to ΛCDM. As in previous
figure two different regimes are considered for computing the PDF. Upper panels θs = 1′, 4′
correspond to the nonlinear calculations where we have assumed a hierarchical ansatz for
the correlation hierarchy. For the lower panels perturbative results are used to construct the
PDF at smoothing angles θs = 15′ and 60′. Various curves correspond to various quintessence
models. In each panel solid line represents the ΛCDM model, short dashed line represent
-- 23 --
wX = −1 + z model, and long dashed line represents wX = −1 + 2z/3 model.
1.2
1.1
1
0.9
1.2
1.1
1
0.9
0.8
-2
-1.5
-1
-0.5
0.8
-2
-1.5
-1
-0.5
Fig. 6. -- Indicator of the deviations of dark energy models from the fiducial ΛCDM model,
1 + ǫ, as function of constant dark energy equation of state wX , for smoothing angle θ0 = 1′
and 15′. The arrows indicate the 1 + ǫ values for dark energy models with time-varying
equation of state, wq(z) = −1 + z (long arrows), and wq(z) = −1 + 2z/3 (short arrows).
800
600
400
200
-- 24 --
800
600
400
200
0
-2
-1.5
-1
-0.5
0
-2
-1.5
-1
-0.5
Fig. 7. -- The skewness S3, for the same models as in Fig.6. The line and arrow types are
the same.
In the left panel we use Hyper-Extended Perturbation theory to compute the
convergence skewness, whereas in the right panel (where larger smoothing angular scales are
considered) perturbative results are adopted.
|
astro-ph/9807246 | 1 | 9807 | 1998-07-23T20:11:34 | A Cluster of Low-Redshift Lyman-alpha Clouds toward PKS 2155-304. I. Limits on Metals and D/H | [
"astro-ph"
] | We report observations from the Hubble Space Telescope (HST) and the VLA on the galactic environment, metallicity, and D/H in strong low-redshift Lya absorption systems toward the bright BL Lac object PKS 2155-304. GHRS/G160M spectra at 20 km/s resolution show 14 Lya absorbers, 6 clustered at cz = 16,100-18,500 km/s. ORFEUS claimed LyC absorption at z = 0.056 with N(HI) = (2-5)x10^16 cm^-2, while our Lya data suggest N(HI) = (3-10)x10^14 cm^-2. Higher columns are possible if the Lya line core at 17,000 +/- 50 km/s contains narrow HI components. We identify the Lya cluster with a group of five HI galaxies offset by 400-800 kpc from the sightline. The two strongest absorption features cover the same velocity range as the HI emission in the two galaxies closest to the line of sight. If the Lya is associated with these galaxies, they must have huge halos of highly turbulent, mostly ionized gas. The Lya absorption could also arise from an extended sheet of intragroup gas, or from smaller primordial clouds and halos of dwarf galaxies. We see no absorption from SiIII 1206, CIV 1548, or DI Lya. Photoionization models yield limits of (Si/H) < 0.003 solar, (C/H) < 0.005 solar, (D/H) < 2.8x10^-4 (4 sigma) if N(HI) = 2x10^16 cm^-2. The limits increase to 0.023 solar and D/H < 2.8x10^-3 if N(HI) = 2x10^15 cm^-2. The data suggest that the IGM in this group has not been enriched to the levels suggested by X-ray studies of intracluster gas and that these absorbers could be primordial gas clouds. | astro-ph | astro-ph |
A CLUSTER OF LOW-REDSHIFT LYMAN-α CLOUDS TOWARD
PKS 2155-304. I. LIMITS ON METALS AND D/H
J. MICHAEL SHULL, STEVEN V. PENTON, JOHN T. STOCKE,
AND MARK L. GIROUX
Center for Astrophysics and Space Astronomy,
Department of Astrophysical and Planetary Sciences,
University of Colorado, Campus Box 389, Boulder, CO 80309
Electronic mail: [email protected], [email protected],
[email protected], [email protected]
J. H. VAN GORKOM AND YONG-HAN LEE
Astronomy Department, Columbia University,
538 W. 120th St., New York, NY 10027
Electronic mail: [email protected]
CHRIS CARILLI
National Radio Astronomy Observatory,
P.O. Box O, Socorro, NM 87801
Electronic mail: [email protected]
ABSTRACT
We report observations from the Hubble Space Telescope (HST) and the VLA on
the galactic environment, metallicity, and D/H in strong low-redshift Lyα absorption
systems toward the bright BL Lac object PKS 2155-304. These studies are intended
to clarify the origin and chemical evolution of gas at large distances from galaxies.
With GHRS/G160M data at ∼ 20 km s−1 resolution, we detect a total of 14 Lyα
absorbers, six of them clustered between cz = 16, 100 and 18,500 km s−1. Although
ORFEUS studies claimed Lyman continuum (Lyc) absorption at z ≈ 0.056 with
N(H I) = (2 − 5) × 1016 cm−2, the Lyα data suggest a range, N(H I) = (3 − 10) × 1014
cm−2. Even higher columns, needed for consistency with the ORFEUS Lyc results, are
possible if the Lyα line core at 17, 000 ± 50 km s−1 contains narrow H I components.
We identify the Lyα cluster with a group of five H I galaxies offset by (400 − 800)h−1
75
kpc from the sightline. The two strongest absorption features cover the same velocity
range as the H I emission in the two galaxies closest to the line of sight. If the Lyα is
associated with these galaxies, they must have huge halos (400 − 500h−1
75 kpc) of highly
turbulent, mostly ionized gas. The Lyα absorption could also arise from an extended
sheet of intragroup gas, or from smaller primordial clouds and halos of dwarf galaxies.
We see no absorption from Si III λ1206, C IV λ1548, or deuterium Lyα at the
expected positions of the strongest Lyα absorbers. Photoionization models yield (4σ)
-- 2 --
limits of (Si/H) ≤ 0.003(Si/H)⊙, (C/H) ≤ 0.005(C/H)⊙, and (D/H) ≤ 2.8 × 10−4 if
N(H I) has the ORFEUS value of 2 × 1016 cm−2. The limits increase to 0.023 solar
metallicity and D/H ≤ 2.8 × 10−3 if N(H I) is only 2 × 1015 cm−2. These limits can
be improved with further studies by HST/STIS and measurements of the Lyc and
higher Lyman series absorption by FUSE. However, the current data suggest that the
intergalactic gas in this group has not been enriched to the levels suggested by X-ray
studies of intracluster gas. Because of their low metallicity and large distance from
galaxies, these absorbers could be primordial gas clouds.
Subject headings: cosmology: observations -- galaxies: abundances -- ISM: H I --
quasars: absorption lines
-- 3 --
1. INTRODUCTION
The origin of the low-redshift Lyα absorption clouds discovered by the Hubble Space Telescope
(Bahcall et al. 1991; Morris et al. 1991; Stocke et al. 1995) remains a mystery, made more
tantalizing by the possibility that they may contain significant mass. Are they remnants of the
high-redshift Lyα forest, or are they associated with past episodes of star formation, galactic
outflows, and galaxy interactions? Based on their frequency, estimated size, and an ionization
correction, they appear to contain a substantial amount (Ωb ≥ 0.003h−1
75 ) of dark baryons (Shull et
al. 1996; Shull 1997). There are also strong indications that some of these absorbers are associated
with galaxies (Lanzetta et al. 1995; Stocke et al. 1995; van Gorkom et al. 1996), although the
distances to the nearest bright galaxies are often quite large (Morris et al. 1993; Stocke et al.
1995; Shull et al. 1996; Grogin & Geller 1998).
From numerical models, it now appears that the high-z Lyα forest is part of a complicated
gaseous structure, formed by the gravitational fluctuations of dark-matter potentials (Cen et
al. 1994; Hernquist et al. 1996; Zhang et al. 1997) and at least partially "polluted" by heavy
element nucleosynthesis during the epoch of galaxy formation. At high redshift, heavy elements
(C IV, Si IV, C II) have been detected at ∼ 10−2.5 times solar metallicity in 50 -- 75% of the Lyα
forest clouds above 1014.5 cm−2 column density (Cowie et al. 1995; Tytler 1995). From X-ray
measurements of Fe-line strengths in galaxy clusters, Mushotzky & Loewenstein (1997) suggest
that intragroup gas might be enriched to levels of 10% solar metallicity at z > 0.4. However,
extrapolating these ideas to low redshift or to all Lyα clouds is difficult. Although a few metal-line
absorption systems have been detected at moderate redshift, no low-redshift Lyα-only absorber
has yet been examined for the presence of heavy elements, which would indicate contamination
by star formation. To investigate this possibility in the Lyα forest requires finding absorbers of
sufficient column density to detect metals [N(H I) > 1016 cm−2] and located appropriately distant
(D > 200 kpc) from neighboring galaxies to suggest primordial gas.
We now believe to have found an ideal target and absorbers, based on our newly analyzed
ultraviolet spectra of PKS 2155-304 from the Hubble Space Telescope (HST) and 21-cm emission
images from the Very Large Array (VLA). Toward this bright, variable (V = 13.1 − 13.7) BL Lac
object, a strong Lyα line at cz ≈ 17, 000 km s−1 was initially identified by the IUE satellite
(Maraschi et al. 1988). This and other Lyα lines were confirmed by low-resolution HST spectra
with the Faint Object Spectrograph (FOS/G130H) (Allen et al. 1993) and with the Goddard
High Resolution Spectrograph (GHRS/G140L) (Bruhweiler et al. 1993). The absorption was
resolved by HST into at least two and possibly three systems. Using fits to Lyman-continuum
(Lyc) absorption near 960 -- 970 A in low-resolution ORFEUS spectra, Appenzeller et al. (1995)
claimed that the 17,000 km s−1 absorbers had a combined column N(H I) ≈ (2 − 5) × 1016 cm−2.
From our HST/GHRS measurements of the corresponding Lyα absorption, we estimate a range,
N(H I) = (3 − 10) × 1014 cm−2. Later in this paper, we will assess the accuracy of the H I column
densities associated with the Lyc absorption and discuss the conditions under which the ORFEUS
and HST measurements could be reconciled.
-- 4 --
Previously, our group observed the PKS 2155-304 field with the VLA (van Gorkom et al.
1996) and found evidence for three H I galaxies corresponding to Lyα absorbers at cz = 5100,
16,500, and 17,000 km s−1. We associated the 17,000 km s−1 absorbers with a small group of
galaxies offset from the PKS 2155-304 sightline by (400 − 800)h−1
75 kpc for a Hubble constant
H0 = (75 km s−1 Mpc−1)h75. In this paper, we combine new HST and VLA observations with
theoretical interpretation of the PKS 2155-304 sightline, which is unusual in the large number
of strong Lyα absorbers and their association with four large galaxies located within ∼ 1 Mpc.
This number of galaxies would be high by chance, based upon the observed two-point correlation
function. The average space density of galaxies in this local region, ngal ∼ 1 Mpc−3, is ∼ 100 times
the large-scale average density of L∗ galaxies (Marzke, Huchra & Geller 1994). This suggests the
presence of a small group that has recently turned around from the Hubble flow, much like our
Local Group. These galaxies probably have not undergone significant mergers, but the presence of
strong, broad Lyα absorption from extended gas may indicate some dynamical interaction.
In § 2.1, we describe new HST/GHRS observations at 20 km s−1 resolution between 1258
and 1293 A (cz = 10, 440 − 19, 060 km s−1). We detect 7 Lyα absorbers with equivalent widths
ranging from 68 to 467 mA. We also reanalyze an archival GHRS/G160M spectrum at 1223 --
1258 A (cz = 1809 − 10, 446 km s−1) and identify 7 definite Lyα absorbers (21 -- 201 mA). In § 2.2,
we present new VLA studies of H I emission in the 2155-304 field from cz = 16, 283 − 17, 571
km s−1, a velocity range that includes four Lyα absorption systems. In § 3, we discuss the results
of these studies. The new VLA data provide accurate correspondence with the Lyα absorbers,
suggesting that the absorption might arise from extended halos or intragroup gas over a region 1
Mpc in diameter that could total 1011 − 1012 M⊙ in mass, if bound. The HST data also allow us
to set limits on the metallicity of the strongest absorbers from the absence of Si III λ1206 and
C IV λ1548, and they provide a limit on D/H from residual (D I) Lyα absorption in the shortward
wing of the strongest absorber. In § 4 we summarize our conclusions and give suggestions for
further study. Somewhat contrary to the kinematic evidence, the low metallicities, [Si/H] ≤ 0.003
solar and [C/H] ≤ 0.005 solar, would suggest that some primordial gas may still reside amidst the
large-scale filaments of galaxies.
2. OBSERVATIONS
2.1. New HST/GHRS Spectra
The target, PKS 2155-304, lies at redshift z = 0.116 or cz = 34, 775 km s−1 (Falomo,
Pesce & Treves 1993). It was observed by HST during Cycle 6 on October 5, 1996, using the
GHRS with the G160M grating and post-COSTAR optics. The continuum flux near 1280 A was
(8.7 ± 0.3) × 10−14 ergs cm−2 s−1 A−1, about 70% of the median flux for this object over the
past 15 years of observations in the IUEAGN database (Penton, Shull & Edelson 1998). Over the
interval 1258 -- 1293 A, we obtained a resolution of 4.5 km s−1 per 0.018 A quarter-stepped pixel or
-- 5 --
∼ 20 km s−1 per resolution element.
The wavelength scale was determined by assuming that the Galactic interstellar S II
absorption features at 1250.584, 1253.811, and 1259.519 A lie at zero velocity in the local
standard of rest (LSR). Based on the Bell Labs 21-cm survey (Stark et al. 1992), the dominant
H I absorption in this direction lies at VLSR ≈ Vhel, within the accuracy (±2 km s−1) of the
observations. This correction in the wavelength scale in the new data (Fig. 1) was +0.029 A,
and that in the archival data (Fig. 2) was −0.001 A. Our observations correspond to redshifted
velocities cz = 10, 440 − 19, 060 km s−1 (z = 0.035 − 0.064) in the Lyα line. According to
previous convention, we quote heliocentric velocities, cz, and compute expected line positions from
λ = λ0(1 + z). We do not make relativistic corrections to the velocities.
Figure 1 shows our new GHRS/G160M data (1258 -- 1293 A) with S/N ≈ 20. Figure 2 shows
our reanalysis of an unpublished archival GHRS/G160M spectrum with S/N ≈ 34 (1223 -- 1258
A) taken with pre-COSTAR optics. The 17,000 km s−1 Lyα absorbers are the strongest of the
low-redshift (z < 0.1) absorbers found in HST searches other than associated absorbers with
zabs ≈ zem. Because of the exceptional brightness of the background source (∼ 13 mag) and
the high H I column densities (≥ 1016 cm−2), these absorbers are ideally suited for measuring
heavy-element abundances in the low-z forest. As of this writing, our group has studied ∼ 100
low-redshift Lyα absorbers toward 11 bright quasars and Seyfert galaxies (Stocke et al. 1995;
Shull et al. 1996; Shull 1997) with more absorbers under analysis (Penton et al. 1998). Our goal
is to understand the physical structure of these clouds and their possible connections with galaxy
halos, large-scale structure, and voids. Our data reduction method was described in these earlier
papers. Complete results on the seven targets in our HST Cycle 6 observations, together with
calibration, are described in Penton et al. (1998).
The GHRS spectra were taken through the 2′′ large science aperture, using the standard
quarter-diode sub-stepping pattern to yield pixels of 0.018 A in FP-split mode. We recalibrated our
spectra using IRAF/STSDAS/CALHRS and the final GHRS reference files (Sherbert & Hulbert
1997) with polynomial background subraction. One noteworthy point concerns our treatment
of the HST/GHRS error vectors, which affect how we gauge the statistical significance of weak
features. We found that the error vectors produced by the IRAF/STSDAS program specalign in
the HST Data Handbook do not agree with those obtained by standard propagation of errors from
individual subexposures. We have therefore chosen to perform our own error propagation and
spectral coaddition using the IDL software package. Spectral coaddition was weighted by exposure
time on a quarter-stepped, pixel-by-pixel basis, removing blemished pixels. This procedure gives
some pixels less exposure time than others, but it prevents photocathode blemishes from being
erronously flagged as absorption features. In some cases, this procedure results in no exposure
time being available for some pixels; straight lines between the last known good flux values are
then used in Figures 1 and 2.
The significance of the detected lines (in σ) is defined as the integrated S/N per resolution
-- 6 --
Fig. 1. -- Top: Smoothed HST/GHRS (G160M) spectrum of PKS 2155-304 shows multiple Lyα
absorption systems, with several strong absorbers at cz ≈ 17, 000 km s−1. Because of smoothing, the
absorbers do not reach zero flux; error vector (1σ) is shown at base. Small arrows near error vector
show deleted photocathode blemishes. Lines at 1259 -- 1261 A are Galactic interstellar absorption;
the weak line at 1259.869 A is Si II in a high-velocity cloud at V ≈ −132 km s−1. Upper limits on
Si III λ1206.50 absorption at 1274.7 A and 1275.2 A (see arrows) correspond to [Si/H] < 0.003 solar
abundance (4σ). Bottom: Six Lyα absorbers (cz = 16, 120 − 18, 590 km s−1) including strong
features near 1281 and 1285 A, estimated to have total N(H I) = (2 − 5) × 1016 cm−2 from Lyc
absorption (Appenzeller et al. 1995).
-- 7 --
Fig. 2. -- Reanalyzed archival GHRS/G160M spectrum from 1223 -- 1258 A shows 7 Lyα lines and
Galactic interstellar lines of S II (1250.584, 1253.811 A) and Mg II (1239.925, 1240.395 A). The Lyα
feature near 1236 A consists of two components separated by 122 km s−1 at comparable velocity to
an H I galaxy at cz ≈ 5100 km s−1 (van Gorkom et al. 1996). We identify the blended absorbers
at 1238.426 and 1238.744 A as Lyα, despite their proximity to Galactic N V (see text).
-- 8 --
element of the fitted absorption feature. The status of "definite absorber" is given to features
greater than 4σ significance (Penton et al. 1998). Table 1 lists the identified absorption features
and their equivalent widths. We see the expected Galactic interstellar lines of S II λ1259.519 and
Si II λ1260.422, as well as a high-velocity cloud in Si II at VLSR = −132 km s−1, which was also
seen in C IV at VLSR = −141 ± 9 km s−1 by Sembach et al. (1998). We also confirm the strong
intergalactic Lyα absorbers near 17,000 km s−1. A blowup of the 1278 -- 1292 A region shows
that many of the Lyα lines have velocity substructure, which we model as separate Gaussian
components. We included these additional components to reduce the χ2 per degree of freedom
whenever the resulting components were above 4σ significance. In no case did we allow more than
two components per blended absorption feature, although we occasionally tested for three.
In addition to the definite Lyα absorbers, we found a number of features at the 3 − 4σ level
that might be verified with future STIS data. Examples of these "possible features" occur at
1229.0, 1243.2, 1249.5, 1264.8, 1286.45, and 1290.58 A. The local continua are uncertain near these
features, and some of the features are too broad (≥ 120 km s−1) to be discrete Lyα absorbers by
our criteria. In these cases, we have decided to be conservative and not classify them as definite.
We also analyzed an archival spectrum (Fig. 2) taken with the GHRS/G160M using
pre-COSTAR optics. This spectrum was reduced and analyzed in a manner identical to that in
Figure 1, with the wavelength scale zero point set to the LSR using the Galactic interstellar S II
lines at 1250.584 and 1253.811 A. Between 1223 and 1258 A, we detect 7 definite Lyα lines and
three Galactic interstellar lines: the two S II resonance lines and the weak Mg II doublet.
The strong Lyα feature at 1236 A is easily detected and is resolved into two subcomponents
separated by 122 km s−1. This pair of lines, at cz = 4999 and 5121 km s−1, lies at the same
velocity as the H I detected galaxy at 21h 57m 04s, −30◦ 25.5 just off the eastern edge of Figure 3
and discussed in detail by van Gorkom et al. (1996). This sub-L∗ galaxy is located ∼ 300h−1
75 kpc
from the PKS 2155-304 sightline. The absorption line at 1238.744 A lies 0.077 A blueward of the
expected position of Galactic N V λ1238.821. We identify this line as Lyα because of the slight
wavelength offset and what would be an unusual strength for N V. Based on the absence of the
weaker line of the N V doublet at 1242.804 A, we believe that N V λ1238.821 contributes at most
30% (at the 4σ level) to the equivalent width of the Lyα feature at 1238.744 A.
The PKS 2155-304 sightline appears to be unusual, both in the number of Lyα absorbers and
in their strength. In total, from 1223 -- 1293 A, we identify 14 Lyα absorbers with significance
greater than 4σ. With a correction for regions blocked by Galactic S II and Si II, these 14 Lyα
absorbers correspond to an uncorrected frequency dN /dz = 250 above 21 mA (NHI ≥ 1012.6
cm−2), for NHI ≥ 1013 cm−2. This frequency is somewhat higher than the mean value, to similar
absorption strength, in sightlines studied with the GHRS during cycles 2 and 4 (Shull et al. 1996).
Probably the most unusual aspect of this sightline, however, is the large number of strong Lyα
absorbers near 17,000 km s−1. The strong correspondence in recession velocity between these
Lyα absorbers and the surrounding galaxies argues that these clouds are not ejected from the
-- 9 --
BL Lac object. Rather, they appear to be intervening clouds at distances given by the Hubble
law, d ≈ (227 Mpc)(V /17, 000 km s−1)h−1
75 , as we assume throughout this paper.
2.2. Deep VLA H I Imaging
Our previous results (van Gorkom et al. 1996) showed that two of the strongest Lyα absorbers
found toward PKS 2155-304 at 17,100 km s−1 are located within a loose group of galaxies. Three
galaxies were detected in H I, two at the velocities of the absorbers, but no individual galaxy
could be identified as being associated with the absorber. The high column densities observed
in the absorbers suggested that there may be an extended component of intergalactic neutral
gas. Column densities N(H I) < 1019 cm−2 are rarely detected in emission (van Gorkom 1993),
and common lore suggests that it may be ionized by the intergalactic UV background. The Lyα
results suggested that this group may be the ideal place to look for diffuse H I at very low column
densities. We therefore reobserved the group in hope of detecting a low surface brightness diffuse
H I component.
The observations were made in 1996 May with the VLA in the 1 km (D) array with an
extended north arm (3 km) to compensate for the low declination of the source. The total
integration time was 40 hrs spread over 8 different runs. All observations were centered at the radio
position of PKS 2155-304, 21h 55m 58.30s, −30◦ 27′ 54.4′′ (B1950). We used a total bandwidth
of 6.25 MHz centered at 16,880 km s−1; the usable velocity range is about 1300 km s−1. On-line
Hanning smoothing was employed, after which every other channel was discarded, leaving a set of
31 independent channels and resulting in a velocity resolution of 46 km s−1. The BL Lac object is
a radio continuum source with a variable flux density. We measured a flux density of 0.45 Jy at 1.4
GHz. Extreme care was taken to properly calibrate the bandpass. Every 25 minutes a bandpass
calibration was done, and the bandpass solution for these individual scans was interpolated to do
the correction. As a result, our observations are limited by noise rather than by spectral dynamic
range, which is better than 3000:1.
The U-V data for each of the 8 days were calibrated independently and inspected for
interference and residual calibration errors. Subsequently, the U-V data of all runs were combined.
The continuum was subtracted by making a linear fit in frequency to the calibrated complex
visibilities of the line free channels (2-5 and 23-25). The resulting data were clipped at a level
of 0.7 Jy to remove man-made and solar interference. Images were made using a taper that
compromised between optimal surface brightness sensitivity and maximal sidelobe suppression
(using the task IMAGR in AIPS, with robustness factor 1), resulting in a synthesized beam
of 54.4′′ × 38.6′′. The rms noise in the channel images is 0.15 mJy beam−1. The instrumental
parameters are summarized in Table 2.
To search for H I, we imaged the entire primary beam (1◦ × 1◦). To calculate H I masses, we
use the luminosity distance (Sandage 1975) assuming that the group is at z = 0.057 and using
-- 10 --
Fig. 3. -- An overlay of the total H I emission (contours) toward PKS 2155-304 in the velocity range
(16,283 -- 17,571 km s−1) on an image of the digitized POSS. Five galaxies are detected in H I at
projected distances from the sightline toward PKS 2155-304 of (400 -- 785)h−1
75 kpc. Contours levels
are 1.65, 3.3, 6.6, 13.2, 19.8, 26.4, 33.0, 39.6 ×1019 cm−2. The image has been corrected for the
primary beam response. Projected distances from the sight line are indicated in h−1
75 kpc. For each
of the galaxies, the velocity range and approximate major axis of the H I emission are indicated.
Note that the galaxy closest to the sightline covers in H I emission exactly the same velocity range
as the broad Lyα absorption seen near 17,000 km s−1. The galaxy to the southwest covers about
the same velocity range as the absorption feature near 16,200 km s−1.
-- 11 --
H0 = 75h75 km s−1 Mpc−1 and q0 = 0.5. Our 6 σ H I mass limit in the center of the field is
(5 × 108 M⊙)h−2
75 . At full resolution, the column density sensitivity is 2 × 1019 cm−2 (5σ over
57 kpc × 34 kpc × 46 km s−1). The data were smoothed spatially and in velocity down to a
resolution 2′ × 3′, resulting in a detection limit of 4 × 1018 cm−2 (5 σ over 126 kpc × 189 kpc
× 92 km s−1). Owing to a lack of short spacings, the observations are much less sensitive to
H I emission that is completely smooth on scales larger than 15′ in a single 46 km s−1 velocity
channel. These values are valid for the center of the field. Farther from the center, they have
to be corrected for the change in primary beam response. The primary beam pattern is roughly
gaussian with a FWHM of 30′, and the limits are a factor two worse 15′ from the field center.
The properties of the five galaxies detected in H I are summarized in Table 3 and shown in
Figure 3. Perhaps the most interesting result of these observations is what we do not find. We
do not detect a faint intergalactic component in H I, nor do we detect any diffuse extended H I
associated with the galaxies. Compared to van Gorkom et al. (1996), our new data are of much
better quality and only show more clearly that the H I emission is confined to the galaxies at
column densities well above 1019 cm−2. In addition to the 3 galaxies detected in the previous
observations, we detect to the southwest an S0 galaxy cataloged in the APM catalog (Loveday
1996) with a velocity range partly outside our band. We see H I to the northwest at 16,375
km s−1, moving toward the center of the galaxy at lower velocities. The lowest velocity where we
have usable, though very noisy data, is at 16237 km s−1. Since this emission peaks only slightly
to the northwest of the center, we infer that the systemic velocity of the galaxy must be about
16200 km s−1. This galaxy was not detected in our previous observations, because in those data
that velocity range was seriously affected by interference.
The other new detection is a tiny dwarf galaxy to the east at 17,156 km s−1. In the current
data, the galaxy only shows up in one channel, but at the 8 σ level. Going back to our previous
data, we found the dwarf just above the noise in our oldest data set (and at the edge of the band
there). Those data were taken with 11 km s−1 velocity resolution, and the emission seems to
cover about 60 km s−1. An overlay of just the dwarf on the digitized POSS is shown in Figure 4.
A hint of some faint light (mB = 22 ± 0.5) can be seen on the blue POSS image, but it is clear
from the image that similar dwarfs could easily be missed optically. In H I the dwarf is just above
our detection limit. Although somewhat brighter galaxies with larger H I masses could have been
detected closer to the sightline, in fact none was seen (Fig. 3).
Are we getting any closer to identifying galaxies associated with the Lyα absorption? There
are 3 galaxies with H I emission in the range 16,900 -- 17,100 km s−1, the velocity of the deepest
Lyα absorption. Thus, in the first instance, it is not obvious that the absorption would be
associated with any one of the galaxies. In our previous work (van Gorkom et al. 1996) we
noticed a curious phenomenon: in all cases, the Lyα absorption was at the systemic velocity of
the closest galaxy. In the current, much improved, Lyα data we notice something even more
curious: the two strong features at 17,100 km s−1 and at 16,200 km s−1 cover exactly the same
velocity range as the H I emission in the two nearest galaxies. In Figure 3 we show the total H I
-- 12 --
5
6
7
8
42
40
RIGHT ASCENSION (B1950)
38
36
34
)
0
5
9
1
B
(
I
N
O
T
A
N
L
C
E
D
I
4
-30 29 45
30 00
15
30
45
31 00
15
30
45
21 56 44
Fig. 4. -- Overlay on the POSS of the dwarf galaxy to the east of PKS 2155-304 sightline at
17,156 km s−1. Contours are same as in Fig. 3. This galaxy has an estimated blue magnitude
mB = 22±0.5. At the observed redshift distance, 230h−1
75 Mpc, its absolute magnitude is MB ≈ −15,
or L ≈ 0.014L∗
B .
-- 13 --
emission contours overlaid on an optical image (greyscale), and we also indicate for each galaxy
the sense of rotation and velocity extent of the H I emission. As in our previous data, there is no
evidence for Lyα corotating with the H I disks. The broad widths of the Lyα absorbers in the
current data is intriguing. It seems unlikely that the two strong absorbers are associated with
their nearest galaxies only. If the width of the Lyα lines were to reflect the potential of the nearest
galaxies, each would have a halo characterized by a flat rotation curve extending out to 400 and
500 kpc respectively. Perhaps more likely, the widths could indicate that the clouds probe a group
potential. In that scenario we may be viewing two groups along the line of sight, one centered at
17,100 km s−1, the other at 16,200 km s−1. The detection of 3 galaxies with systemic velocities
close to 17,100 km s−1 makes it plausible that there is indeed a loose group at this velocity. Our
velocity coverage does not extend to sufficiently low velocities to make any statements about the
presence of a group at 16,200 km s−1.
3. RESULTS AND DISCUSSION
3.1. Extended Halo and Intragroup Gas
The combined HST and VLA data suggest that the 17,000 km s−1 Lyα absorption lines arise
in a group environment. Between 16,100 and 18,100 km s−1, we observe 6 Lyα absorbers and four
large (VLA) galaxies at similar velocities. The velocity centroid and radial velocity dispersion of
the four galaxies are hVgali = 16, 853 km s−1 and σgal = 325 km s−1. The equivalent statistics
for the 6 Lyα absorbers are hVLyi = 17, 070 km s−1 and σLy = 740 km s−1. The fact that the
strongest Lyα absorbers (1284.484 A and 1285.097 A) lie close to hVLyi suggests that the center of
the gravitational potential well may be centered at V ≈ 17, 000 km s−1. Before we interpret the
larger velocity dispersion of Lyα lines, it is worth noting that this galaxy group was observed in
H I emission with the VLA over a more limited velocity range (16,283 -- 17,571 km s−1) than the
Lyα lines.
The simplest interpretation of the data is that the absorbing gas arises in intragroup material
stripped or blown out of the gas-rich galaxies. However, there is no direct evidence for this
interpretation in the H I emission data. Nor are galaxy H I halos expected to be this dense at
such enormous radii (Dove & Shull 1994). Combining the 740 km s−1 velocity dispersion of the
6 definite Lyα absorbers with the mean projected separation, hRi ≈ 1.0 Mpc between the four
galaxies, we obtain a virial mass estimate,
Mvir ≈
3σ2
r hRi
2G
≈ (2 × 1014 M⊙)h−1
75 .
(1)
The estimated velocity dispersion and size/mass scale are those of a modest group, less than the
size of the Virgo cluster. The mean distance of these galaxies at hVgali = 16, 853 km s−1 is 225h−1
75
Mpc or (m − M ) = 36.8. The break in the standard luminosity function is L∗ = (9 × 109L⊙)h−2
75
or M∗ = −19.4 + 5 log h75 (Marzke et al. 1994) in the Zwicky (blue) magnitude system. Thus, an
-- 14 --
L∗ galaxy in this group would lie at mB ≈ 17.4. Using two slightly brighter galaxies in the field
for calibration of the SRC-J plate magnitudes (Lauberts & Valentijn 1989), we have estimated B0
magnitudes for the five galaxies around PKS 2155-304. In terms of L∗, they are: north (0.3L∗),
east (0.9L∗), south (1.4L∗), southwest (3.0L∗), and dwarf (0.014L∗). Thus, three of the H I
galaxies have luminosities essentially at or above L∗.
Although drawing statistical inferences from only four galaxies is risky, we have attempted
to estimate the possible number of fainter galaxies, both observationally and theoretically. In a
CFHT image (Wurtz et al. 1997) centered on PKS 2155-304, of radius 160h−1
km s−1, we found five galaxies in excess of the background to a limit of mB ≈ 21, about 3.5
magnitudes below L∗. Over the larger field of 600h−1
75 kpc radius, defined by the four H I galaxies,
this corresponds to ∼ 60 excess galaxies. However, this photometry does not constrain the redshift
of these excess galaxies; most, if not all, could lie at the redshift, zem = 0.116, of PKS 2155-304
(Falomo et al. 1993). Since BL Lac objects are routinely found in poor clusters (Wurtz et al.
1997), these five excess galaxies are likely to be associated with PKS 2155-304.
75 kpc at cz = 17, 000
If we integrate a Schechter luminosity function, φ(L) = φ∗(L/L∗)−1 exp(−L/L∗), down to
0.04L∗ (mB ≈ 21), we would expect between 3 and 10 times more galaxies, compared to those
above 0.3L∗ and 0.9L∗, respectively. Scaling from the observed H I galaxies above those limits,
we would expect to find between 10 and 30 galaxies associated with this group, down to 0.04L∗.
Therefore our preliminary optical survey is consistent with these expectations. However, since our
radio H I survey detected only one dwarf galaxy in the field, gas-rich dwarfs may be scarce in this
group. We are conducting a spectroscopic survey of the PKS 2155-304 field down to mB ≈ 19 to
test our hypothesis that the large VLA galaxies are accompanied by smaller galaxies, some closer
to the PKS 2155-304 sightline.
We also investigated the possibility that hot gas associated with the strong absorbers at
17,100 km s−1 might be responsible for the 600 eV absorption feature seem by the objective
grating spectrometer on the Einstein Observatory (Canizares & Kruper 1984) and by the BBXRT
spectrograph (Madejski et al. 1998). This possibility would only work if the X-ray absorption arose
from an O V K-edge (626 eV rest frame, 592 eV observed) or an O VIII Kα line (653 eV rest frame,
618 eV observed). We can rule out the O V hypothesis (T ≈ 2 × 105 K) because it would require
an unreasonably large hydrogen column, NH = (1.7 × 1022 cm−2)ζ −1
O , for oxygen metallicity ζO, in
order to produce the observed optical depth, τ ≈ 2.5. The resulting gas densities would be quite
high, across a 300 kpc slab, and the cooling time would be less than 105 yrs. Line absorption by
O VIII would be more feasible, requiring a column density NH ≈ (1.3 × 1019 cm−2)ζ −1
O . However,
neither of these cluster hot-gas models would explain the observed broad (∼ 30, 000 km s−1)
absorption widths. We conclude that the 600 eV absorption must be produced elsewhere, probably
within PKS 2155-304.
Because the four bright galaxies are H I-rich spirals, it is also possible that these galaxies are
not bound, based upon recent spectroscopic work on similar galaxy groups (Mulchaey & Zabludoff
-- 15 --
1998; Zabludoff & Mulchaey 1998). Even if they are bound, these galaxies may not have virialized
or closely interacted. If this is the case, the Lyα clouds may also be non-virialized, and our
estimate of Mvir would be too large. This interpretation suggests that these clouds are portions
of the gaseous filament out of which these spiral galaxies formed. If these clouds and galaxies are
portions of a bound group, then we would expect the clouds to have metallicities of 0.1 -- 0.3 solar,
similar to the stripped and virialized gas in rich clusters and elliptical-rich groups (Mushotzky
& Loewenstein 1997). In the unbound case, not only would X-ray emission be unlikely, but the
metallicity of the gas would be substantially lower than 10% solar (see § 3.2).
In our earlier papers on the environments of the low-z absorbers (Shull et al. 1996; van
Gorkom et al. 1996), we estimated a space density, φ0 ≈ (0.7 Mpc−3)R−2
with N(H I) ≥ 1013 cm−2 and characteristic size scale (100 kpc)R100, in order to explain their
frequency per unit redshift. This space density is ∼ 40 times larger than that of L∗ galaxies, but
comparable to that of dwarf galaxies with L ≈ 0.01L∗, suggesting a possible connection. However,
the H I absorption cross sections of these dwarfs are uncertain, and it is unclear whether the
extended gas was stripped out, blown out, or existed primordially.
100h75, of low-z Lyα clouds
The main results of our comparison between the Lyα absorption and the 21-cm emission
can be summarized as follows: (1) 21-cm emission has been detected in this region, spread over
1 Mpc of sky and 1300 km s−1 of velocity; (2) the large distances to the nearest bright galaxies
(400-800h−1
75 kpc); and (3) the agreement in velocity (within ±100 km s−1) between several of the
Lyα absorbers and the H I galaxies to the south, southwest, and north. The proper interpretation
of the Lyα absorbers toward PKS 2155-304 hinges on several geometrical issues. Does the Lyα
absorption occur in a smoothly distributed layer, 200-600 kpc in depth, or in smaller, denser
clumps? Some evidence for the latter interpretation comes from the fact that the Lyα absorption
occurs in discrete systems in velocity space. Less certain is whether the absorbers are kinematically
associated with the large galaxies. Although a single sight line through the region is not necessarily
typical, the VLA observations suggest that the absorbing gas is distributed across a region ∼ 1
Mpc in diameter. As a first approximation, we will model the absorption as a homogeneous slab.
We then explore more complex distributions, either in intergalactic clouds or in dwarf-galaxy
halos.
In the homogeneous approximation, the total mass of gas in the vicinity of the cluster of
galaxies around PKS 2155-304 can be estimated by assuming a slab of radius 500 kpc, depth
D = 400 kpc, mean column density N(H I) = 2 × 1016 cm−2, exposed to an ionizing radiation field
with specific intensity I0 = 10−23 ergs cm−2 s−1 Hz−1 sr−1 at 1 Ryd and spectral slope αb = 1.8,
as described in § 3.2. (The BL Lac, PKS 2155-304 cannot dominate the ionizing flux incident on
these clouds, unless its actual redshift is much less than the value, z = 0.116, from Falomo et al.
1993.) If I0 and D are constant, then N(H I) ∝ nHI ∝ n2
H, so that total density, nH, and gas mass,
Mgas, scale as [N(H I)]1/2. For the parameters above, the mean densities are hnHIi = 2.2 × 10−8
-- 16 --
cm−3 and hnHi = 4.4 × 10−5 cm−3, and the total gaseous mass, including helium, is
Mgas = (4 × 1011 M⊙)(cid:20) R
500 kpc(cid:21)2(cid:20)
N(H I)
2 × 1016 cm−2(cid:21)1/2
.
(2)
The gas mass could therefore approach 1% of the 1014 M⊙ required to bind the group, a relatively
small gas fraction compared to the gas found in clusters and elliptical-rich groups. Likewise, the
luminous mass in these galaxies is small. If the gas is clumped into denser parcels with the same
total covering factor, the neutral fraction would increase and the total gas mass would decrease
inversely with the characteristic depth of the absorbers.
Let us now consider inhomogeneous models for the gas distribution around PKS 2155-304,
involving dwarf galaxies or primordial gas filaments. Let the total projected area of the gas be
L × L with depth D, where L ≈ 1 Mpc and D ≈ 400 kpc. Assume that this volume is filled by
an ensemble of Ncl clouds, each of radius r ≈ 40 kpc, with a velocity dispersion 750 km s−1. The
cloud filling factors in area and volume are, fA = πNcl(r/L)2 and fV = (4π/3)Ncl(r/L)3(L/D),
where fA could exceed 1 if clouds overlapped in projected area. Using the given parameters, their
ratio is (fA/fV ) = (4/3)(r/D) ≈ 0.13. Since we detected many absorbers along the sightline, it is
likely that fA ≥ 1, which suggests that fV ≈ 0.3 − 0.5. A consequence of this model is that these
"clouds" should collide with one other on a crossing time tcr= (0.6 Mpc)/(750 km s−1) < 109 yr.
Therefore, if this group is bound, we would expect considerable gas stripping, with the possibility
of shocks, hot gas (T ≈ 5 × 106 K), and tidal plumes. There is no evidence for these effects,
although there has been no search for extended soft X-ray emission or O VI absorption from this
region.
If the group of galaxies and clouds is not bound, then this cloud ensemble could either be
primordial gas which never participated in galaxy formation or gaseous halos of luminous or
dwarf galaxies. Based upon an extensive survey of quasar fields, for which ultraviolet spectra
were obtained by the HST Key Project Team, Lanzetta et al. (1995, hereafter L95) identified
∼ 1/3 of large equivalent width (Wλ ≥ 0.3 A) Lyα absorption lines with bright galaxies at impact
parameters ≤ 210h−1
75 kpc away from the sightline. The three strongest close pairs of Lyα clouds in
Table 1 would not be resolved into multiple components at the 1 A resolution of the Key Project
spectra. While all three cloud complexes are easily strong enough to be included in the L95 study,
none is as close to a bright galaxy as those in the L95 survey: 300h−1
absorber, 560h−1
absorber. Our H I detection survey in this field (van Gorkom et al. 1996 and § 2.2 herein) reaches
depths at least comparable to the L95 survey (0.1 -- 0.3 L∗). Thus, the bright galaxy halos in this
field may be anomalously larger than those found by L95 in other fields, and also substantially
larger than expected theoretically (Dove & Shull 1994). However, we strongly suspect that bright
galaxy halos cannot account for these clouds.
75 kpc for the 5100 km s−1
75 kpc for the 17,000 km s−1
75 kpc for the 16,300 km s−1 absorber, and 400h−1
Are there undetected dwarf galaxies at the Lyα cloud redshifts closer to the PKS 2155-304
sightline? If so, then dwarf-galaxy halos could be responsible for these absorptions as extrapolated
-- 17 --
by L95 and others. Chen et al. (1998) found a best-fit Lyα halo size that scales with galaxy
luminosity as rh ∝ L0.35, a somewhat stronger dependence than the rh ∝ L0.15 found for Mg II
absorbing halos (Steidel 1995). With the Chen et al. scaling relation, the two weaker absorption
complexes (5100 and 16,200 km s−1) would be consistent with the L95 and Chen et al. observations
if any fainter galaxies (L ≤ 0.1L∗) were found closer to PKS 2155-304 at those redshifts. However,
the 17,000 km s−1 absorber is so strong that a dwarf galaxy would need to be very close to the
sightline to satisfy the Chen et al. halo-size relationship. Specifically, we would require impact
parameters ρ < 67h−1
75 kpc for L ≈ 0.01L∗. The only bright
galaxy sufficiently near to PKS 2155-304 to satisfy the first condition is galaxy G4 of Falomo et
al. (1993), which has L ≈ 0.2L∗ and ρ ≈ 80h−1
75 kpc. However, G4 has z = 0.117, the same as
PKS 2155-304. While there are some closer galaxies, their numbers are modest: 1 -- 3 excess
galaxies to L ≤ 0.01L∗ based on the independent data from Wurtz et al. (1997) and Falomo et
al. (1993). The only two galaxies with spectroscopy in hand are galaxies 1 and 2 of Falomo et
al. (1993) at z = 0.117, located 4 arcsec E and 25 arcsec SE of the BL Lac, respectively. At this
point, there is no dwarf galaxy candidate for any of these three strong absorption complexes.
75 kpc for L ≈ 0.1L∗ and ρ < 33h−1
Thus, although unlikely, the dwarf-galaxy halo hypothesis remains possible. We detected one
75 from the PKS 2155-304 sightline, and we may find other faint
faint dwarf (MB ≈ −15) at 565h−1
galaxies at the redshifts of the three strong absorption complexes. However, there is no direct
evidence in favor of the dwarf-halo, and the primordial cloud hypothesis must be taken seriously
for all three Lyα absorbers.
3.2. Metallicity Limits
We can also use our spectra to set limits on the metallicity of the strongest Lyα absorbers.
As seen in Figure 1, metal lines should be searched for in three strong absorption systems,
corresponding to Lyα lines at λ1 = 1281.393 A (z1 = 0.054063), λ2 = 1284.484 A (z2 = 0.056605),
and λ3 = 1285.097 A (z3 = 0.057110). In the limited wavelength range of our GHRS/G160M
spectrum, the only expected strong metal line is the Si III λ1206.500 resonance line. No Si III
absorption is present at the 1271.73 A location of system 1. Although we see a hint of Si III
absorption at the expected positions [1274.79 A and 1275.40 A, see Fig. 1] corresponding to
systems 2 and 3, we treat this absorption as upper limits. Using the unsmoothed data to detect
weak, unresolved features, we obtain a formal 4σ error on equivalent width of 23 mA (rest-frame
22 mA).
To convert the observed Si III/H I to limits on abundances (Si/H), we must make an
ionization correction based on photoionization conditions in the absorbers. Appenzeller et al.
(1995) claimed to detect weak Lyc absorption with the far-UV spectrograph aboard ORFEUS.
The Lyα clouds at z = 0.054 − 0.057 were estimated to have a combined column density N(H I) =
(2 − 5) × 1016 cm−2. We have derived the expected column densities of metal ions by modeling
the strong absorbers as slabs with N(H I) = 2 × 1016 cm−2 and total depth ranging from 5 to
-- 18 --
800 kpc, comparable to or less than the offset distance from the nearest galaxies. The slabs are
assumed to be illuminated on both sides by an ionizing spectrum with specific intensity at 1 Ryd
of I0 = 10−23 ergs cm−2 s−1 Hz−1 sr−1, consistent with a recent calculation by Shull et al.
(1998)
that gives I0 = (1.1 ± 0.5) × 10−23, based on low-z Seyfert galaxies and a new IGM opacity model.
The spectral index of the ionizing background is taken as αs = 1.8 ± 0.1 (Zheng et al. 1997).
As a "standard model", we assume a homogeneous slab, of depth 400 kpc and 0.003 solar
metallicity. The photoionization equilibrium is computed with the model CLOUDY Version 90.03
(Ferland 1996). The results are given in Table 4. In general, the C III λ977 and C IV λ1548 lines
are the best tracers of metals, while Si III λ1206 is a factor of 5 -- 10 weaker, assuming relative
solar abundances of (Si/H)⊙ = 3.55 × 10−5 and (C/H)⊙ = 3.63 × 10−4 (Grevesse & Anders 1989).
However, the C III line lies shortward of the HST band, and its observation must await the launch
of the FUSE satellite in early 1999. For unsaturated absorption, the predicted equivalent widths
can be written:
Wλ(C IV) = (73 mA)(cid:20)
Wλ(Si III) = (22 mA)(cid:20)
N(H I)
2 × 1016 cm−2(cid:21)(cid:20) [C/H]
0.003(cid:21) ,
2 × 1016 cm−2(cid:21)(cid:20) [Si/H]
0.003 (cid:21) .
N(H I)
(3)
(4)
Following the convention of our earlier Lyα work, we treat "definite absorbers" as those with
4σ significance. The observed 23 mA (4σ) limit on Si III λ1206 (rest equivalent width 22 mA)
corresponds to N(Si III) ≤ 1.0 × 1012 cm−2. If we scale the column density of the strong absorption
components to N(H I)= 2 × 1016 cm−2, the (4σ) upper limit on metallicity can be written,
H(cid:21) ≤ (0.003)" 2 × 1016 cm−2
(cid:20) Si
N(H I)
#(cid:18) Si
H(cid:19)⊙
.
(5)
Another limit on metallicity comes from C IV λ1548.195. From their low-resolution
(GHRS/G140L) spectrum, Bruhweiler et al. (1993) quote a (2σ) upper limit of 110 mA for C IV
λ1548. We recalibrated the pre-COSTAR GHRS/G140L spectrum using IRAF/STSDAS/CALHRS
with the final calibration files. As before, subexposure coaddition was performed using our own
IDL routines, which merge the subexposures by exposure time with photocathode blemishes
removed (Fig. 5). We find a 4σ limit of Wλ ≤ 134 mA (127 mA rest frame) or N(C IV) ≤ 3 × 1013
cm−2 for a linear curve of growth. This yields a metallicity limit,
H(cid:21) ≤ (0.005)" 2 × 1016 cm−2
(cid:20) C
N(H I)
#(cid:18) C
H(cid:19)⊙
.
(6)
Figures 6 and 7 illustrate how these metallicities depend on the two key parameters of the
photoionization models: the mean hydrogen density hnHi and the spectral slope αb of the ionizing
background. The observed column density, N(H I), and the assumed cloud depth, D, determine
-- 19 --
Fig. 5. -- GHRS/G140L data of PKS 2155-304 (Bruhweiler et al. 1993) reduced with our software
and GHRS final calibration files to show the spectral region for C IV λ1548.195 absorption
corresponding to the three strong Lyα absorbers at 1281.393, 1284.484, and 1285.097 A (see Table
4). The signal-to-noise ratio is 30 per 0.029 A pixel. The C IV upper limits are each 134 mA (4σ)
in the observed frame.
-- 20 --
Fig. 6. -- Contours of column density, N(Si III), corresponding to N(H I) = 2×1016 cm−2 and 0.003
solar metallicity. The axes represent the assumed spectral slope of the ionizing background spectra
(1.7 ≤ αb ≤ 1.9) and cloud depths (5 ≤ D ≤ 1000 kpc). Hatched lines mark regions disallowed
by the 4σ non-detection of Si III, while asterisks mark the first viable solutions, D ≥ 300 kpc
and D ≤ 40 kpc, consistent with the standard value αb = 1.8. Large-D solutions correspond to
absorption from a homogeneous slab, while small-D solutions could arise in denser galactic halos
along the sightline.
-- 21 --
Fig. 7. -- Same as Fig. 6, showing contours of column density, N(C IV), corresponding to N(H I)
= 2 × 1016 cm−2 and 0.003 solar metallicity. Hatched lines mark regions disallowed by the 4σ
non-detection of C IV, while the asterisk marks the first viable solutions, D ≤ 400 kpc, consistent
with αb = 1.8. Although the combined Si III and C IV limits allow metal-bearing clouds of size
200 -- 400 kpc, there is more parameter space at sizes less than 40 kpc.
-- 22 --
the mean neutral density,
hnHIi = (1.6 × 10−8 cm−3)(cid:20)
N(H I)
2 × 1016 cm−2(cid:21)(cid:20) 400 kpc
D (cid:21) .
(7)
The neutral fraction, hnHI/nHi, is set by photoionization equilibrium and is proportional to the
ionization parameter, U ∝ J0/nH, as are the other metal-ion column densities. The precise
behavior of the ratios (Si III/H I) and (C IV/H I) depend on the range of U where these particular
ions peak in fractional abundance (Donahue & Shull 1991). The spectral slope, αb, affects the
ionization rates above the relevant ionization thresholds (16.34, 33.46, 45.13 eV for Si II, III, IV;
24.38, 47.87, 64.48 eV for C II, III, IV). For a fixed radiation intensity, J0, decreasing the assumed
cloud depth D results in a higher hnHi, smaller ionization parameter U , larger hydrogen neutral
fraction, and lower N(C IV) and N(Si III). For αb = 1.8 ± 0.1 and D = 400 ± 200 kpc, the predicted
column densities are:
N(Si III) = (1.0+0.2
N(C IV) = (1.8+0.7
−0.2 × 1012 cm−2)(cid:18) [Si/H]
0.003 (cid:19)
−0.5 × 1013 cm−2)(cid:18) [C/H]
0.003(cid:19) .
(8)
(9)
The topology of C IV and Si III differ, as seen in the excluded regions of Figures 6 and 7.
For denser absorbers, the ionization equilibrium shifts to lower ionization parameter (U ) in which
case C IV/H I decreases (Fig. 7) while Si III/H I remains about the same (Fig. 5). For Si III,
the ionization solutions can be double-valued, since Si III changes more strongly with U than H I
(Donahue & Shull 1991). At sufficiently low and sufficiently high ionization parameter, silicon lies
in other states (Si II or Si IV). As a result, the H I absorption toward PKS 2155-304 could still
contain metals at 0.01 solar and arising either in a homogeneous slab, of depth D =200-400 kpc,
or in denser parcels such as 10 -- 40 kpc halos of dwarf galaxies or intergalactic clouds.
Thus, it is easier to hide metals from detection in the C IV lines than in Si III. The observed
limit on Si III absorption with N(H I) = 2 × 1016 cm−2 and 0.003 metallicity formally allows
solutions with D > 400 kpc or D < 40 kpc. Limits on C IV absorption allow D < 400 kpc.
The combined C IV and Si III limits, together with the numerous discrete Lyα absorbers in
velocity space, suggest that the gas may be inhomogeneous on scales of tens of kpc. However, it is
puzzling how clumped absorbers could maintain a high density contrast when these clumps would
be expected to collide on a billion-year time scale. Such collisions would also shock-heat the gas
possibly making C IV lines more detectable. Filamentary sheets, with large aspect ratios, provide
a more plausible scenario consistent with these constraints.
Before concluding this section, it is worth examining the accuracy of the measurements of
N(H I). Our assumed value, N(H I) = 2 × 1016 cm−2, comes from the claimed detection of a Lyc
depression in ORFEUS data (Appenzeller et al. 1995). These data suggest a flux discontinuity
shortward of 970 A, corresponding to the z = 0.054 − 0.057 Lyα absorbers, but the Lyman break
-- 23 --
is not definitive, owing to the moderate (0.5 A) spectral resolution, low signal-to-noise (S/N = 15),
and uncertainties in defining and extrapolating the true continuum between 920 and 1100 A. For
similar reasons, these data are not sufficiently accurate to determine N(H I) from a curve-of-growth
analysis of the higher Lyman series, Lyβ -- Lyδ.
Because of Lyα line saturation in our HST/GHRS data, we cannot confirm the large H I
columns implied by the Lyc absorption. However, we can provide both a firm lower limit and give
reasonable estimates from the strongest Lyα features at 1284 -- 1285 A. The optical depth of the
Lyα line per unit velocity is τ (v) = (πe2/mec)f λN (v), where N (v) is the H I column density per
unit velocity. We may integrate this, for f = 0.4164 and λ = 1215.67 A, to obtain,
Ntot = (7.45 × 1011 cm−2)Z τ (v) dv ,
(10)
where v is measured in km s−1. In our post-COSTAR data, with S/N ≈ 20, this formula provides
only a minimum column density, since we cannot distinguish optical depths τ (v) > 3. Flux
calibration and background subtraction are difficult with the GHRS one-dimensional detectors.
An integration of the data (Fig. 8) yields values Ntot ranging from (3 − 10) × 1014 cm−2. Values
of 2 × 1016 cm−2 needed for consistency with the ORFEUS Lyc measurement would require
unresolved narrow components in the line core.
Although we have some skepticism about adopting the (2 × 1016 cm−2) column density from
ORFEUS, it is worth remembering that line saturation and multiple velocity components make
column density measurement extremely difficult from just one line. A prime example of this
difficulty comes from the Lyα absorbers seen toward 3C 273 at cz ∼ 1000 km s−1 and cz ∼ 1600
km s−1 (Weymann et al. 1995). Recent ORFEUS measurements of these components (Hurwitz
et al. 1998) find Lyβ absorption equivalent widths larger by factors of 1.5 and 2.4 than values
predicted by Lyα Voigt profile fits of HST/GHRS data (log N = 14.19 and b = 40.7 km s−1 for the
1000 km s−1 cloud, and log N = 14.22, b = 34.2 km s−1 for the 1600 km s−1 cloud). By analyzing
both Lyα and Lyβ, Hurwitz et al. (1998) derive a best-fit of log N = 15.8 and b = 17 km s−1.
They conclude that, if the H I absorption arises in a single cloud, its column density is higher by
at least a factor of 4 compared to the value of Weymann et al. (1995).
Thus, it could be that the strong, saturated Lyα lines towards PKS 2155-304 have columns
well above 1015 cm−2 and even as high as 2 × 1016 cm−2. This could arise if narrow H I components
are hidden within the line core at 17, 000 ± 50 km s−1. For example, 20 components, each with
b ≈ 15 km s−1 and N(H I) ≈ 1015 cm−2, spread stochastically over ∼ 100 km s−1 would reproduce
the Lyα and Lyc data. Each component would have central optical depth τ0 ≈ 50 and an (isolated)
equivalent width of 240 mA. Saturation and velocity overlap would allow their accumulated Lyα
absorption to give the observed ∼ 0.8 A, while the Lyc optical depth could reach values τc ≈ 0.1.
Without better data on the Lyc absorption edges or the higher Lyman series lines, we
cannot verify the Lyα absorption (2 × 1016 cm−2) in the strong absorber at cz = 16, 970 km s−1.
Confirmation of the Lyman limit and higher Lyman series will await our planned studies of
-- 24 --
Fig. 8. -- Blow-up of the Lyα absorption cluster at cz ≈ 17, 000 km s−1. Left: Lyα absorption line
vs. wavelength and heliocentric velocity, cz, showing the continuum flux (solid line at top) and
1σ background (dashed line at bottom). Right: optical depth, τ (v) of the same absorption line,
computed from left panel, assuming that the continuum lies at the dashed line. We ignore negative
fluxes and values τ > 25. Owing to uncertainties in flux calibration and background subtraction,
values τ > 3 are unreliable. The integral of this absorption yields N(H I) = (3 − 10) × 1014 cm−2
(the higher value for right panel shown here). Larger columns, consistent with the ORFEUS LyC
estimates, N(H I) = (2 − 5) × 1016 cm−2, would require unresolved narrow components in the line
core.
-- 25 --
PKS 2155-304 following the scheduled Feb. 1999 launch of the Far Ultraviolet Spectroscopic
Explorer (FUSE). With FUSE, if N(H I) is as large as 2 × 1016 cm−2, the Lyman edge at 963.36 A
will have τc ≈ 0.13, and many higher Lyman lines will be detectable. Even if we adopt the
minimum values, N(H I) ≥ 4 × 1014 cm−2 and τ (Lyα) ≥ 3 in the line core, we should be able to
detect higher Lyman lines with τ (Lyβ) ≥ 0.48, τ (Lyγ) ≥ 0.17, and τ (Lyδ) ≥ 0.08. It would also
be helpful to obtain a better Lyα spectrum of the 1270-1290 A region with HST/STIS, whose
two-dimensional array detectors will improve the background subtraction and flux calibration.
The metal-line searches can also be improved with HST, using STIS and COS (see § 3.4).
If N(H I) is smaller than the value (2 × 1016 cm−2) assumed in Figures 5 and 6, our limits on
metallicity increase. For homogeneous clouds, of constant density, several quantities follow from
simple scaling relations of N(H I). First, our inferred value of D/H scales inversely with N(H I),
and would therefore go up. Second, one would find a tradeoff of cloud depth (D) and metallicity
([C/H] or [Si/H]). If N(H I) were decreased by a factor 10, the metallicity limits would rise by
the same factor. If D is held constant, a simple scaling of the metallicity with N(H I) is not
possible. As we discussed earlier, the relative ionization fractions of H I, C IV, and Si III all vary
with ionization parameter, U . If we retain our estimate of the radiation background and assume
D = 400 kpc, reducing N(H I) by a factor of 10 changes the metallicity limits to C/H< 0.023 and
Si/H< 0.14 times solar. For the range in ionization parameter −1.9 < log U < −1.3 appropriate to
these assumptions about cloud density and radiation field, the greater sensitivity to U of the Si III
fraction causes the Si III constraints on metallicity to weaken much more rapidly with decreasing
N(H I). Regardless of the exact value of N(H I), the most sensitive metal lines remain upper limits.
Therefore, one must take seiously the possibility that these clouds are primordial.
3.3. Limits on Deuterium
A similar analysis can be done using the strong absorbers to search for weak deuterium Lyα
(1215.339 A) features in the shortward wings of H I Lyα (1215.667 A). No deuterium feature
is seen in the wings of the H I system 1 (H I at 1281.393 A, D I at 1281.044 A) to a level of
Wλ ≤ 20 mA. In system 2 (H I at 1284.484 A, D I at 1284.133 A) the limit is Wλ ≤ 30 mA.
For a linear curve of growth, these equivalent widths translate to N(D I) ≤ 3.7 × 1012 cm−2 and
5.5 × 1012 cm−2, respectively. If we attribute column densities N(H I) of 1 × 1016 cm−2 and
2 × 1016 cm−2 to systems 1 and 2, respectively, these limits give a deuterium abundance limits of,
H(cid:19) ≤ (3.7 × 10−4)" 1 × 1016 cm−2
(cid:18) D
H(cid:19) ≤ (2.8 × 10−4)" 2 × 1016 cm−2
(cid:18) D
N(H I)
N(H I)
# (system 1)
# (system 2)
(11)
(12)
These limits can be improved significantly with higher-precision data on the D I line (with
-- 26 --
HST/STIS) and with better H I column densities from higher Lyman-series lines (with the FUSE
spectrograph). At the current level of accuracy, neither limit provides a strong constraint on Big
Bang nucleosynthesis (Schramm & Turner 1998). The interstellar medium value has been reported
in the range D/H = (1.6 ± 0.09) × 10−5 (Linsky et al. 1993, 1995). However, a controversy still
exists over the high-redshift D/H observed in QSO absorption systems. A recent determination
of D/H in two QSO Lyman-limit systems (Burles & Tytler 1997) gives a "low value", D/H
= (3.4 ± 0.25) × 10−5, while Songaila and Cowie (cf. Songaila 1997) quote "high values" with
a range D/H = (0.4 − 1.5) × 10−4. The high values of D/H would obviously require a large
destruction rate of deuterium through star formation. Thus, it would be helpful to obtain a
detection of deuterium in the strong Lyα systems at z = 0.054 − 0.057 toward PKS 2155-304.
Here, the measured metallicity is much less than solar values, implying little astration of D. We
expect that future HST and FUSE observations can make a factor of 3 improvement in sensitivity,
which could set a limit on D/H < 1 × 10−4.
3.4. Future Searches for Metal Absorption
Our photoionization models demonstrate that, for clouds with ≥ 0.001 solar metallicity, it
should be possible, with better HST/STIS data, to detect the 1548.2 A and 1550.8 A resonance
lines of C IV. We may also be able to detect C II λ1335, N V λ1238, Si IV λ1394, and Si III
λ1206. With the FUSE satellite, we intend to search for C III λ977, C II λ1036, and O VI
λ1032. The C III line should be detectable down to well below 0.001 solar metallicity. The C II
lines might be present if the gas is at higher density (lower U ). The O VI line is weaker in the
standard model (D = 400 kpc, αb = 1.8), but it might be enhanced by a number of effects: a hard
photoionizing radiation field above 114 eV, a collisionally ionized component due to hot gas, or
oxygen abundance enhancement by massive-star nucleosynthesis. We probably will not detect O I
λ1302, which is weak, even at 3% solar metallicity, because the O I abundance is tied by charge
exchange to the ratios (H II/H I) and (O II/O).
Because of the O VI ionization correction, the strength of the λ1032 absorption line depends
both on the shape of the ionizing spectrum and on the cloud depth. For example, if αb = 1.4
(D = 400 kpc) the O VI column density increases by over a factor of 4, while if αb = 1.8
(D = 600 kpc) it increases by a factor of 2. The total H I column in the absorbers could be
higher than the assumed value, 2 × 1016 cm−2. In addition, the Si and O abundances could be
enhanced (Songaila & Cowie 1996; Giroux & Shull 1997) relative to C in regions dominated by
"prompt nucleosynthesis" from massive-star supernovae. Thus, if Si/C > 2 × (Si/C)⊙ and O/C
> 2 × (O/C)⊙, we may be able to detect lines of O VI, Si II, Si III, and Si IV.
As shown by Donahue & Shull (1991), the ratios of column densities from multiple ion states
of the same element (e.g., Si II/III/IV and C II/IV) can be modeled to provide diagnostics of the
intensity and spectrum of the 1 -- 4 Ryd ionizing radiation field. Detections of these ions would
allow us to discriminate between an extragalactic radiation field due to AGN or due to O-stars
from starburst galaxies.
-- 27 --
4. CONCLUSIONS
The sightline to PKS 2155-304 is unique among the AGN studied thus far for low-redshift Lyα
absorbers. It has the highest frequency of absorbers, and they are identified with a concentration
of bright galaxies. The Lyα absorbers at cz ≈ 17, 000 km s−1 may be an extreme example of the
previously known association of Lyα clouds with the extended halos of galaxies (Lanzetta et al.
1995; Stocke et al. 1995). Of greater interest is the enormous inferred gaseous extent; the nearest
bright galaxies lie (400 − 800)h−1
75 off the sightline. These offsets are so large that they are unlikely
to represent equilibrium gaseous disks; the orbital times at such distances would be enormous. We
have interpreted these absorbers as large sheets of intragroup gas, or as smaller primordial clouds
and halos of dwarf galaxies. The clumpiness of the Lyα absorption in velocity space suggests that
some spatial structure is present. However, as discussed in § 3.1, a medium that is clumpy in both
space and velocity would be subject to disruptive collisions.
Our major conclusions are therefore:
• The metallicity limits for the Lyα absorbers give (Si/H) and (C/H) less than 0.003 solar
if N(H I) = 2 × 1016 cm−2 (the value from ORFEUS). The greatest uncertainty in these
metallicities is the precise value of N(H I). We will use the FUSE satellite to measure the
H I columns through Lyman limit absorption and higher Lyman series lines. We hope to see
discrete absorption edges at 961.04, 963.36, and 963.82 A.
• Our metallicity limit, [Si/H] < 0.003 solar contradicts the hypothesis that these clouds are
metal-enriched intragroup gas. Mushotzky & Loewenstein (1997) suggest that intracluster
gas might be enriched early (at z > 0.4) to levels of 10% solar metallicity. Evidently, the
intergalactic gas around PKS 2155-304 has not been enriched to the levels observed in X-ray
emitting intracluster gas.
• If the "small group" hypothesis is correct, a search for X-rays with AXAF imagers would
test whether any hot gas has been produced by tidal effects or stripping.
• Future spectral observations with HST can provide even better limits on D/H, Si III, and
C IV than those reported here. With FUSE, we will be able to search for the expected
strong C III λ977 line, the weak O VI λ1032, as well as absorption lines of S II, C II, S II,
and N II which may be present if the absorption occurs in denser halos.
• The low metallicity limits and the large projected distances from the nearest galaxies suggest
that these clouds could be primordial.
-- 28 --
This work was based 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 and on observations made with NRAO's Very Large Array. The NRAO is operated by
Associated Universities, Inc. under a cooperative agreement with the National Science Foundation.
We thank Michael Fall and Richard Mushotzky for helpful discussions on the evolution of
intracluster gas. This work was supported by HST Guest Observer grant GO-06593.01-95A and
by the Astrophysical Theory Program (NASA grant NAGW-766 and NSF grant AST96-17073) to
the University of Colorado and by an NSF grant (AST96-17177) to Columbia University.
-- 29 --
Table 1
HST Absorption Featuresa in PKS 2155-304
Wavelength
(A)
1226.362 ± 0.054
1226.990 ± 0.062
1232.032 ± 0.045
1235.941 ± 0.050
1236.436 ± 0.100
1238.426 ± 0.099
1238.744 ± 0.065
1239.817 ± 0.040
1240.411 ± 0.052
1250.588 ± 0.015
1253.807 ± 0.011
1259.519 ± 0.007
1259.869 ± 0.019
1260.398 ± 0.055
1270.802 ± 0.014
1281.393 ± 0.006
1281.867 ± 0.063
1284.484 ± 0.009
1285.097 ± 0.013
1287.515 ± 0.010
1288.976 ± 0.020
Velocityb
(km s−1)
2637 ± 13
2789 ± 15
4035 ± 11
4999 ± 12
5121 ± 24
5612 ± 24
5690 ± 16
Galactic
Galactic
Galactic
Galactic
Galactic
Galactic
Galactic
13, 596 ± 4
16, 208 ± 2
16, 325 ± 16
16, 970 ± 2
17, 121 ± 3
17, 717 ± 3
18, 078 ± 5
Rest EW Significancec
ID
(mA)
42 ± 24
36 ± 18
21 ± 11
129 ± 7
201 ± 8
21 ± 10
85 ± 22
33 ± 12
20 ± 11
80 ± 11
127 ± 12
132 ± 16
75 ± 20
467 ± 81
102 ± 17
345 ± 22
68 ± 30
448 ± 22
363 ± 24
141 ± 16
100 ± 19
(σ)
9.0
8.0
4.2
29
45
4.8
20
7.5
4.5
19
28
15
8.8
53
12
45
9.2
61
48
19
13
Lyα
Lyα
Lyα
Lyα
Lyα
Lyα
Lyα
Mg II
Mg II
S II
S II
S II
Si II (HVC)d
Si II+Fe II
Lyα
Lyα
Lyα
Lyα
Lyα
Lyα
Lyα
a Top group of lines measured from previously unpublished, pre-COSTAR GHRS/G160M
spectrum (Fig. 2). Bottom group is from our post-COSTAR, GHRS/G160M spectrum (Fig. 1).
We use global continuum fits and quote rest-frame equivalent widths (EW).
b LSR velocity (cz), which is the same as heliocentric velocity for this sightline to within ±1
km s−1. Wavelength scales have been aligned assuming that the Galactic S II lines lie at VLSR.
This procedure gives offsets of +0.029 A (new data) and −0.001 A (pre-COSTAR data).
c Significance of the line (in σ) is defined as the integrated S/N per resolution element of the fitted
absorption feature.
d High velocity cloud was detected in C IV at VLSR = −141 ± 9 km s−1 (Sembach et al. 1998).
-- 30 --
Table 2
VLA Instrumental Parameters
Date
Configuration
Integration time (hrs)
Central Velocity (km s−1)
Channel Width (kHz)
Channels
Velocity Resolution (km s−1)
Usable Velocity Range (km s−1)
Synthesized beam (arcsec)
rms noise (mJy/beam)
rms column density sensitivity (cm−2)
1996 May
DnC
40
16,880
195
31
46
16,283 - 17,571
54.4 × 36.6
0.15
4 × 1018
Table 3
H I Properties of Detected Galaxies
Name
RA (1950) Dec (1950)
Vhel
Velocity Range
MHI
LB/L∗
(km s−1)
(109 M⊙)
(km s−1)
17,087
16,785
17,179
17,156
2155-3033
F21569-3030
uncataloged
APMBGC 466-053+024
uncataloged
21 55 46.9
21 56 56.0
21 56 03.6
21 55 29.9
21 56 38.5
16,888 - 17,294
-30 33 49.6
16,650 - 16,926
-30 30 43.7
-30 17 08.7
17,064 -17,294
-30 33 52.6 ∼ 16,200 <16,237 -16,375
-30 30 57.4
17,156
7.4
14
4.9
> 3.6
0.8
1.4
0.9
0.3
3.0
0.014
-- 31 --
Table 4
Predicted Column Densitiesa and Equivalent Widths
Ion
Column Density Wλ
(mA)
(cm−2)
λ0
(A)
Observed λi
[λ0(1 + zi)]
C II
C III
C IV
N V
O VI
Si II
Si III
Si IV
1.1 × 1012
2.4 × 1013
1.8 × 1013
1.7 × 1012
1.6 × 1012
6.5 × 1010
1.0 × 1012
5.8 × 1011
2.2
156
73
3.6
2.0
1.0
22
4.5
1334.532
977.020
1548.195
1238.821
1031.926
1260.422
1206.500
1393.755
1406.68, 1410.07, 1410.75
1029.84, 1032.32, 1032.82
1631.90, 1635.83, 1636.61
1305.80, 1308.94, 1309.57
1087.72, 1090.34, 1090.86
1328.56, 1331.77, 1332.40
1271.73, 1274.79, 1275.40
1469.11, 1472.65, 1473.35
a Column densities are computed for cloud with NHI = 2 × 1016 cm−2, D = 400 kpc, and 0.003
solar metallicity, irradiated by QSO ionizing spectrum with αb = 1.8 and I0 = 10−23 ergs cm−2
s−1 Hz−1 sr−1. Observed wavelengths correspond to Lyα components at 1281.393, 1284.484, and
1285.097 A, at redshifts z1 = 0.054063, z2 = 0.056605, and z3 = 0.057110. Equivalent widths are
quoted in the rest frame.
-- 32 --
REFERENCES
Allen, R. G., Smith, P. S., Angel, J. R. P., Miller, B. W., Anderson, S. F., & Margon, B. 1993,
ApJ, 403, 610
Appenzeller, I., Mandel, H., Krautter, J., Bowyer, S., Hurwitz, M., Grewing, M., Kramer, G., &
Kappelmann, N. 1995, ApJ, 439, L33
Bahcall, J. N. et al. 1991, ApJ, 377, L5
Bruhweiler, F. C., Boggess, A., Norman, D. J., Grady, C. A., Urry, C. M., & Kondo, Y. 1993,
ApJ, 409, 199
Burles, S., & Tytler, D. 1997, preprint (astro-ph/9712265)
Canizares, C. R., & Kruper, J. 1984, ApJ, 278, L99
Cen, R., Miralda-Escud´e, J., Ostriker, J. P., & Rauch, M. 1994, ApJ, 437, L9
Chen, H.-W., Lanzetta, K. M., Webb, J. K., & Barcons, X. 1998, ApJ, 498, 77
Cowie, L. L., Songaila, A., Kim, T.-S., & Hu, E. M. 1995, AJ, 109, 1522
Donahue, M., & Shull, J. M. 1991, ApJ, 383, 511
Dove, J. B., & Shull, J. M. 1994, ApJ, 423, 196
Falomo, R., Pesce, J. E., & Treves, A. 1993, ApJ, 411, L63
Ferland, G. 1996, Hazy, University of Kentucky Internal Report (Version 90.03)
Giroux, M. L., & Shull, J. M. 1997, AJ, 113, 1505
Grevesse, N., & Anders, E. 1989, in Cosmic Abundances of Matter, AIP Conf. 183, ed. C. J.
Waddington, 1.
Grogin, N. A., & Geller, M. J. 1998, ApJ, in press (astro-ph/9804326)
Hernquist, L., Katz, N., Weinberg, D. H., & Miralda-Escud´e, J. 1996, ApJ, 457, L51
Hurwitz, M., et al. 1998, ApJ, 500, L61
Lauberts, F., & Valentijn, E. A. 1989, The Surface Photometry Catalog of the ESO-Uppsala
Galaxies, (Garching, ESO)
Linsky, J. L., et al. 1993, ApJ, 402, 694
Linsky, J. L., et al. 1995, ApJ, 451, L335
Loveday, J. 1996, MNRAS, 278, 1025
Madejski, G., et al. 1998, preprint, submitted to ApJ
Maraschi, L., Blades, J. C., Calanchi, C., Tanzi, E. G., & Treves, A. 1988, ApJ, 333, 660
Marzke, R. O., Huchra, J. P., & Geller, M. J. 1994, ApJ, 428, 43
Morris, S., Weymann, R. J., Savage, B., & Gilliland, R. L. 1991, ApJ, 377, L21
-- 33 --
Morris, S., Weymann, R. J., Dressler, A., McCarthy, P. J., Smith, B. A., Terrile, R. J., Giovanelli,
R., & Irwin, M. 1993, ApJ, 419, 524
Mulchaey, J. S., & Zabludoff, A. I. 1998, ApJ, 496, 73
Mushotzky, R., & Loewenstein, M. 1997, ApJ, 481, L63
Penton, S., Shull, J. M., & Edelson, R. 1998, Database of IUE-AGN Ultraviolet Spectra, available
on Website (http://casa.colorado.edu/∼spenton/IUEAGN/FUSE.html)
Penton, S., Stocke, J. T., & Shull, J. M. 1998, in preparation.
Sandage, A. 1975, in Galaxies and the Universe, ed. by A. Sandage, M. Sandage, J. Kristian
(University of Chicago Press, Chicago).
Schramm, D. N., & Turner, M. 1998, Rev. Mod. Phys., 70, 303
Sembach, K. R., Savage, B. D., Lu, L., & Murphy, E. M. 1998, ApJ, in press
Sherbert, L. E., & Hulbert, S. J. 1997, GHRS Instrument Science Report 067, Official Update
(July 1997)
Shull, J. M. 1997, in Structure and Evolution of the IGM from QSO Absorption Lines, ed. P.
Petitjean & S. Charlot, (Paris, Editions Fronti`eres), 101
Shull, J. M., Stocke, J. T., & Penton, S. 1996, AJ, 111, 72
Shull, J. M., Roberts, D., Giroux, M., & Penton, S. 1998, in preparation.
Songaila, A. A. 1997, in Structure and Evolution of the IGM from QSO Absorption Lines, ed. P.
Petitjean & S. Charlot, (Paris, Editions Fronti`eres), 339
Songaila, A., & Cowie, L. L. 1996, AJ, 112, 335
Stark, A. A., Gammie, C. F., Wilson, R. W., Bally, J., Linke, R. A., Heiles, C., & Hurwitz, M.
1992, ApJS, 79, 77
Steidel, C. C., 1995, in QSO Absorption Lines, Proc. of ESO Symposium, ed. G. Meylan
(Heidelberg, Springer), 139
Stocke, J., Shull, J. M., Penton, S., Donahue, M., & Carilli, C. 1995, ApJ, 451, 24
Tytler, D. 1995, in QSO Absorption Lines, Proc. of ESO Symposium, ed. G. Meylan (Heidelberg,
Springer), 289
van Gorkom, J. H. 1993, in The Environment and Evolution of Galaxies, ed. J. M. Shull & H. A.
Thronson, (Dordrecht, Kluwer), 345
van Gorkom, J. H., Carilli, C. L., Stocke, J. T., Perlman, E. S., & Shull, J. M. 1996, AJ, 112, 1397
Weymann, R. J., Rauch, M., Williams, R., Morris, S., & Heap, S. 1995, ApJ, 438, 650
Wurtz, R., Stocke, J. T., Ellingson, E., & Yee, H. K. C. 1997, ApJ, 480, 547
Zabludoff, A. I., & Mulchaey, J. S., 1998, ApJ, 496, 39
Zheng, W., Kriss, G. A., Telfer, R. C., Grimes, J. P., & Davidsen, A. F. 1997, ApJ, 475, 469
-- 34 --
This preprint was prepared with the AAS LATEX macros v4.0.
|
astro-ph/9606038 | 1 | 9606 | 1996-06-06T11:39:56 | Accretion disks in interacting binaries: Simulations of the stream-disk impact | [
"astro-ph"
] | We investigate the impact between the gas stream from the inner Lagrangian point and the accretion disk in interacting binaries, using three dimensional Smooth Particle Hydrodynamics simulations. We find that a significant fraction of the stream material can ricochet off the disk edge and overflow towards smaller radii, and that this generates pronounced non-axisymmetric structure in the absorption column towards the central object. We discuss the implications of our results for observations and time-dependent models of low-mass X-ray binaries, cataclysmic variables and supersoft X-ray sources. | astro-ph | astro-ph | ACCRETION DISKS IN INTERACTING BINARIES:
SIMULATIONS OF THE STREAM-DISK IMPACT
P.J. Armitage1
M. Livio2
ABSTRACT
We investigate the impact between the gas stream from the inner Lagrangian
point and the accretion disk in interacting binaries, using three dimensional Smooth
Particle Hydrodynamics simulations. We find that a significant fraction of the stream
material can ricochet off the disk edge and overflow towards smaller radii, and that this
generates pronounced non-axisymmetric structure in the absorption column towards
the central object. We discuss the implications of our results for observations and
time-dependent models of low-mass X-ray binaries, cataclysmic variables and supersoft
X-ray sources.
ApJ accepted
6
9
9
1
n
u
J
6
1
v
8
3
0
6
0
6
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1Institute of Astronomy, Madingley Road, Cambridge, CB3 0HA, UK
2Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA.
-- 2 --
1.
INTRODUCTION
Accretion disks in interacting binaries are unlikely to have simple axisymmetric structures.
In these systems, the accretion disk is fed with material from a companion star via an accretion
stream that originates at the inner Lagrange point. In this picture, both the gravitational influence
of the companion and the impact of the stream with the disk provide mechanisms that may distort
the shape and vertical structure of the accretion disk. Modeling of these deviations from axial
symmetry requires the use of two and three dimensional hydrodynamic models, and is necessary
in order to interpret a variety of observations of cataclysmic variables (CVs), low-mass X-ray
binaries (LMXBs) and supersoft x-ray sources (SSS).
The X-ray light curves of nearly edge-on LMXBs provide one observational clue to the
existence of non-axisymmetric disk structures. The eclipses are typically broad, with a gradual
ingress that commences well before the phase of first contact (e.g. White et al. 1994). In some
sources ("dipping" sources), most notably X1822-371, there is an actual dip in the flux at an
orbital phase near 0.8. This absorption has been interpreted as arising from a bulge in the height
of the disk rim at that phase (White & Holt 1982; Mason 1989 and references therein), which is
suggestively similar to the location at which the accretion stream would impact the disk edge.
X-ray observations of GK Persei (Hellier & Livio 1994) and EUVE observations of CVs (Warren,
Sirk & Vallerga 1995; Long et al. 1996) provide further evidence for observational signatures of
the accretion stream and its interaction with the disk.
Two dimensional calculations of accretion disk structure have primarily been used to
investigate the formation of the disk and its response to the tidal field of the companion (Lin &
Pringle 1976; Whitehurst 1988a,b; Lubow 1991; Murray 1996). Using a variety of particle-based
numerical schemes, these studies show that for systems with a low mass ratio of secondary to
primary (q <
∼ 0.25), the accretion disk is tidally unstable, leading to the formation of an eccentric
precessing disk. Proper consideration of the effect of the accretion stream on the disk (which acts
to add low specific angular momentum material at the outer edge and so tends to reduce the
disk radius) shows that this modifies the phenomenon but does not alter the general behaviour
(Murray 1996).
Hirose, Osaki & Minishige (1991) carried out simplified three dimensional calculations in
an attempt to examine whether variations in disk scale height caused by the stream impact
could explain the dipping X-ray sources. They found thickenings of the disk at around the
phases expected, if these were the causes of modulation in the X-ray light curves. However, the
difficulties of modeling the entire disk for many dynamical times meant that the resolution of
the stream-disk impact region in this calculation was poor, and the authors commented that an
improved treatment of the hydrodynamics beyond their "sticky particle" scheme was desirable.
Much better resolution of the impact region was achieved in calculations by Livio, Soker & Dgani
(1986) and Dgani, Livio & Soker (1989), but this was achieved at the expense of considering only a
small region surrounding the point where the stream met the disk. Thus, while these calculations
-- 3 --
demonstrated clearly the flow pattern in the collision region, there remained the possibility that
the hydrostatic vertical structure assumed for the disk gas might not be realistic, if the response
of the entire disk to the stream was included. With this caveat, however, the calculations did
demonstrate that stream material could overflow the disk rim, as suggested by the work of Lubow
& Shu (1975, 1976), Lubow (1989) and Frank, King & Lasota (1987).
In the current investigation, we extend these studies by performing three dimensional Smooth
Particle Hydrodynamics (SPH) simulations of the response of the disk to the accretion stream.
We focus on the dynamical effects of the stream-disk collision -- for which the SPH method is
well-suited -- and aim for the highest practicable resolution of the interaction region. We do not
attempt to model the thermal or long-term viscous evolution of the disk, for which purpose other
numerical techniques would be preferable. By ignoring these aspects we are able to improve the
resolution of the stream-disk impact region when compared, for example, to prior 3D simulations
by Lanzafame, Belvedere & Molteni (1994a, 1994b), who were interested primarily in modeling
outburst behaviour over longer timescales than those considered here.
The layout of this paper is as follows. In Section 2 we describe the computational methods
and binary system parameters that we adopt for the simulations. Section 3 presents and interprets
the results of the calculations, while Section 4 summarizes our conclusions and suggests further
observational tests of the models.
2. COMPUTATIONAL APPROACH
2.1. Smooth Particle Hydrodynamics
The simulations described in this paper were performed using a three dimensional SPH code
minimally modified from that described by Benz (1990). SPH is a fully lagrangian, particle-based
hydrodynamics method, that has been applied to a wide variety of problems in astrophysics (see,
e.g. Monaghan 1992), and has been shown to give closely comparable results to hydrodynamics
codes based on alternative numerical techniques (e.g. PPM -- Davies et al. 1993).
For the simulation of accretion disks, SPH has a number of advantages. A Keplerian disk
exhibits a rapid variation of dynamical timescale with radius, which is computationally expensive
if the problem demands that a significant range of radii be modeled. This difficulty can be
partially alleviated in an SPH calculation by assigning each particle its own timestep satisfying
the Courant stability criterion, so that the bulk of the disk material at relatively large radii can
be integrated on a longer timestep. SPH has further merits in facilitating easy tracing of the flow,
and in not wasting resources following the void regions outside and above the accretion disk.
Disadvantages of this method include the variable resolution of an SPH calculation, which is
worst in the low density parts of the flow where the particle density is small. As these regions -- well
-- 4 --
above the mid-plane of the disk -- are precisely those that are most important for determining
the X-ray and EUV absorption, this is a potentially serious problem. We mitigate its effect by
exploiting the freedom that SPH allows in setting particle masses, so that the stream particles
(which on collision with the disk will form the majority of the material high above the disk plane)
have lower masses that those in the disk. More problematic is the fact that our code neglects the
effects of radiation transport and cooling of the gas, so that the only control over the thermal
properties arises via the choice of the equation of state. We adopt an isothermal equation of
state, which corresponds to assuming that the energy generated from viscous and shock heating
is radiated away instantaneously. For a standard α accretion disk (Shakura & Sunyaev 1973),
the thermal timescale, tth ∼ α−1Ω−1, is comparable to the dynamical timescale, tdyn ∼ Ω−1, if
the Shakura-Sunyaev viscosity parameter α is of order unity (Pringle 1981). Arguments based
on the timescales involved in the outbursts in these systems suggest that for CVs and LMXBs
α is typically in the range 0.1 -- 1, so that the isothermal equation of state provides a reasonable
approximation to the much more complicated physics that is actually involved. Caution is
nevertheless required in interpreting our results close to where shocks -- and hence strong and
rapid heating of the gas -- occur in the simulations.
2.2. System Parameters
Having ignored both thermal and viscous effects, the important system parameters that
remain are the mass ratio q = M2/M1, the disk mass Mdisk, and the accretion rate M . For low
mass ratios q <
∼ 0.25, prior calculations show the formation of an eccentric, tidally unstable disk,
which precesses in the rotating frame of the binary (Whitehurst 1988a,b; Lubow 1991). Such mass
ratios are likely to be relevant both to subclasses of dwarf novae (the SU UMa systems), and to
many LMXBs with low-mass (a few tenths M⊙) companions. Conversely, higher mass ratios lead
only to a modestly distorted disk which remains fixed in the Roche frame of the binary.
The disk mass and accretion rate are expected to be significant mainly via the combination
f = M P/Mdisk, where P is the period of the binary. f is then the fraction of the disk mass that
is replenished per orbit, and it measures how strong a perturbation the stream can exert on the
disk. To estimate sensible values for f , we assume a Shakura-Sunyaev disk in which the viscosity
is given by ν = αcsH, where cs is the sound speed and H ≈ cs/Ω is the disk scale height. For such
a disk, the surface density Σ in the steady state is given by,
νΣ =
M
3π
,
(1)
for radii much larger than the inner disk radius. Taking Mdisk ∼ πR2Σ, we then obtain an estimate
for f,
f =
,
(2)
M P
Mdisk ∼ 6πα(cid:18) cs
vK(cid:19)2
-- 5 --
where vK is the Keplerian velocity in the disk and all quantities are evaluated at the disk edge.
For a thin disk cs/vK ≪ 1 (≈ 0.03 for our simulations), so that f is small, typically of order
10−2, even for α of order unity. This is also consistent with the timescales of outbursts (in which
mass accumulated in the disk over an extended time rapidly drains onto the central star), which
generally last ∼ 102P .
In previous calculations the most dramatic effects of the companion on the accretion disk
were found to occur for low mass ratios. We therefore adopt q = 0.2, as appropriate for a
0.3 M⊙ companion to a 1.4 M⊙ primary, and f = 0.025. For the particular value of Mdisk in the
M = 2 × 1017 gs−1, or 3 × 10−9 M⊙ yr−1
calculations, this corresponds to a mass transfer rate of
(run LMXB2). To assist in disentangling the effects of the Roche potential from those caused
solely by the impact of the stream, a second simulation was run with a very low mass secondary
(q effectively zero), and otherwise similar values of disk mass and accretion rate (run LMXB1).
2.3. Disk and Stream Set Up
As we have already remarked, simulating the innermost regions of the disk is computationally
prohibitive. Fortunately it is also unnecessary, since on a dynamical timescale the stream is unable
to influence the disk interior to the stream circularization radius. We therefore set up the disk
as an annulus with an inner edge at approximately the circularization radius and an outer edge
at the tidal truncation radius as given by Papaloizou & Pringle (1977) and Paczy´nski (1977). To
prevent a small fraction of particles from bringing the simulation to a halt by moving too far
inward under the action of viscosity, we add angular momentum to any disk particle straying
inside a radius Rbound from the primary. For the runs described here, Rbound is set at half the
inner radius of the annulus. Over the relatively short timescale of our runs, we observe no effect
of this inner boundary condition on the disk annulus farther out.
Although the effects of physical viscosity are unimportant for these calculations, it is important
that the artificial viscosity in the code, parameterized by αSPH and βSPH (Benz 1990), is sufficient
to prevent particle interpenetration in the shocks that occur when the stream strikes the disk edge.
An isothermal equation of state provides the least resistance to particle interpenetration (as there
is no shock heating to generate additional pressure gradients), and so we adopt the rather high
values αSPH = 2, βSPH = 4. Previous work (Bate 1995) has shown that this should be adequate to
stop particles streaming past each other, and we do not see any such behaviour in our runs. The
internal SPH viscosities can be converted into a physical Shakura-Sunyaev α parameter via the
formula (e.g. Bate 1996),
α ≈ (cid:18) αSPH
20 (cid:19)(cid:18) h
H(cid:19) ,
(3)
where h is the particle smoothing length in the simulation (3D). The β-viscosity will also have
some effect, but for a disk that is resolved in the vertical direction, the contribution to the
-- 6 --
Fig. 1. -- (a) Distribution of particles in z for a thin annulus in a quiescent (no stream) disk. The
dashed line shows a gaussian distribution with width ∆z = 2 × 109 cm. (b) The disk scale height
evaluated in cells for the same quiescent disk, plotted as a function of radius. The dashed line
represents H ∝ R3/2.
kinematic viscosity from the second-order βSPH term will be smaller than that from αSPH (Bate
1996). Using this expression it can be seen that for simulations that are resolved vertically (h < H,
as in our calculations), typical SPH viscosity parameters lead to values of α that are smaller than
those believed to pertain to CV and LMXB accretion disks, and there should be no problems
arising from unreasonably large artificial viscosity in the simulations.
The disk gas is modeled using 30000 equal mass particles that are initially distributed on a
close-packed lattice within 2.5◦ of the disk plane, and which have Keplerian velocities about the
primary. The sound speed is set such that cs/vK = 0.03 at the outer edge of the disk -- this gives
a somewhat thicker disk than is usually assumed, but has the advantage that better resolution
is attained in the vertical direction. The disk reaches a stable hydrostatic equilibrium structure
quickly (in less than one orbit), after which the distribution of particles with z (perpendicular to
the disk) for a narrow annulus of disk is as shown in Fig. 1(a). The particles follow a gaussian
distribution, as expected for an isothermal accretion disk, and track the density profile well, up to
the height where the particle number density becomes very low. Calculating the disk scale height
-- 7 --
(defined as H = Σ/ρz=0, where Σ and ρz=0 are evaluated from the SPH calculation) as a function
of radius yields the results shown in Fig. 1(b). For an isothermal disk, H ∝ cs/Ω ∝ R3/2, and this
scaling is shown in the Figure as the dashed line. Over the range (−0.3 < log(R) < 0) where the
disk is resolved vertically this relation is found to be obeyed, with discrepancies at the inner and
outer edges where the particle density is dropping off rapidly.
For run LMXB2 the presence of the companion means that the disk is not being set up with
velocities that are consistent with disk orbits in the binary potential. To allow time for these
initial conditions to be smoothed out, the disk is first evolved for ∼ 10 orbits of the binary before
the stream flow is switched on. During this relaxation period the disk develops a stable but
distorted shape, with a smooth variation of surface density. For the purposes of the present work
this represents well enough the 'background' disk structure onto which the stream flows, though
if, for example, one wished to follow the disk precession in a 3D calculation, a more elaborate disk
set up would certainly be desirable.
For the initial conditions of the stream, we utilise the results of Lubow & Shu (1975, 1976), see
also review by Livio (1994). These show that the width of the stream, W , in both the azimuthal
and vertical directions, is given approximately by,
W ∼
cs
Ωbinary
,
(4)
where cs is here the sound speed at the surface of the mass-losing star and Ωbinary is the binary
angular velocity. This is essentially the same relation as for the disk scale height, and implies that
the stream is much thicker than the pressure scale height of the mass-donating star H∗ (indeed,
W ∼ √H∗R∗).
Numerically, the stream is modeled as a tube of particles stretching outward from the inner
Lagrange point, L1, with radial velocities set to be cs towards the primary. When a particle passes
inward of L1 (towards the primary) it is added to the simulation and evolved under the influence
of gravity and all hydrodynamic forces, while particles outside L1 are simply rotated to follow
the binary orbit. Initially the stream particles are strongly compressed in the radial direction
(by a factor ∼ 10). This ensures that when the stream reaches the disk edge the separations
are comparable in R and in z, as required to model the impact hydrodynamics. We assume
that the stream leaves L1 in hydrostatic equilibrium, and set the vertical density profile to be a
gaussian with width W . This is truncated at 1 scale height, where the scale height is estimated by
extrapolating the numerical results obtained from the disk (Fig. 1), and allowing, where necessary,
for the gravity of the secondary. The compression required in R near L1 means that the stream
there is not in numerical hydrostatic equilibrium, but this is of no consequence, since the stream
flow toward the disk is ballistic. The simulations follow the stream-disk interaction for around
2-3 orbits of the outer disk, requiring 20000 -- 30000 particles to achieve adequate resolution of the
accretion stream.
-- 8 --
Fig. 2. -- Particle distribution in the x − y and x − z planes for simulation LMXB2 (mass ratio
of secondary to primary q = 0.2). This figure is omitted due to size, and is available from
http://www.ast.cam.ac.uk/∼ parm/lmxb.html
3. RESULTS
3.1. Stream Overflow
Fig. 2 shows a snapshot of particle positions during the latter stages of run LMXB2. For
both simulations, a quasi-steady state appearance in the Roche frame is rapidly attained, so that
although the disk mass is slowly increasing throughout, successive snapshots look very similar.
Where the stream reaches the disk, the Figure shows that the oblique collision both shears the
stream and compresses the disk edge. Some stream material is transferring its momentum to the
disk gas and becoming buried in the disk rim. This is similar to the behaviour seen in previous
calculations (e.g. Livio et al. 1986; Rozcyczka & Schwarzberg-Czerny 1987). However, from the
x − z projection it is also clear that some stream gas is overflowing the disk rim as a result of the
collision. This material is being thrown to larger distances above the midplane than the original
stream, and is many scale heights above the disk itself.
To clarify the origin of this overflowing gas, we plot in Fig. 3 density contours in a vertical
plane, aligned along the direction of the accretion stream flow where it reaches the disk edge.
We also show velocity vectors in this plane for the stream gas, calculated by simple averaging
of particle velocities in spheres of radius 2h. Although the disk in general flares towards large
radius (H ∝ R3/2), it can be seen that in the outermost regions of the disk where the surface
density is dropping, the density contours form a wedge-like structure. The stream strikes this
supersonically, and the outer parts ricochet off to greater distances from the disk midplane. This
behavior has been predicted by the analytic Sedov Solution obtained by Livio et al. (1986). The
vertical velocities away from the z = 0 plane that are achieved in this manner, are typically several
times cs, so this gas reaches a maximum z that is many times the disk scale height. Further
downstream the overflowing gas falls back to the disk, achieving higher velocities than on impact,
because now it is closer to the primary where the vertical component of gravity is greater. While
the central core of the stream tends to blunt the wedge, it is continually restored by fresh disk
material that possesses a higher angular velocity than the orbital one.
3.2. Density Slices from the Simulations
Figs. 4 and 5 show the density calculated using the SPH smoothing algorithm for both
simulations. We plot the midplane density, panel (a), and the density at a few scale heights above
-- 9 --
Fig. 3. -- Logarithmic density contours (dashed lines) in a vertical plane aligned with the direction
of stream flow at the location where the stream strikes the disk. Vectors represent the velocity in
this plane of stream material. The z-component of the velocity vectors has been enhanced by a
factor of two for clarity.
the midplane, panel (b). In the latter plot, we have co-added 4 independent time-slices in order to
reduce the Poisson noise due to limited particle number, though the behavior we describe is clearly
seen also in individual snapshots. For both simulations the compression of the disk rim due to the
momentum of the stream impact is evident (the oscillations in radius of the disk, downstream of
the impact, are probably due to remaining anisotropies in particle distribution on impact, and are
not physical). However, except for this, the midplane density plots show little structure beyond
what would be expected given the initial set up of these disks.
At high z, conversely, strong signatures of the stream impact are visible. For LMXB1 a
single region of relatively high density is obtained, beginning at the point where the stream reaches
the disk and persisting until around phase 0.75 (here phases are measured with respect to the
phase at which an eclipse would occur, and increase clockwise in the plots). This peak lies well
inside the outer edge of the disk, and arises from the overflowing stream material discussed earlier.
Although the resolution at this z is limited by relatively small particle numbers, it appears that
the overflowing gas does not form a well-collimated continuation of the accretion stream. Rather
-- 10 --
Fig. 4. -- Density from simulation LMXB1 (secondary of zero mass) in the x− y plane at (a) z = 0,
and (b) z = 0.15.
the impact leads to a broad fan of material ejected to high z as a consequence of the collision.
For the simulation with the full Roche potential, a similar structure generated by the
stream-disk interaction is seen in the density plots. In this instance however, a second broad and
comparatively high density region is seen at phase 0.2, which originates from the eccentricity of
the accretion disk in this simulation. The disk is flared, and so for a given high z slice, the greatest
density contribution is expected to come from material at the largest radius. This leads to a peak
roughly in the direction along which the disk is most extended, though from the simulation the
peak is seen at a somewhat earlier phase, probably as a consequence of the disk gas not reaching
hydrostatic equilibrium and hence lagging in adjusting to the changing vertical gravity. We note
that the magnitude of the second peak is probably overestimated by the simulation, as the SPH
particles at the outer edge of the disk have no neighbours at greater radii and so adopt larger
smoothing lengths to maintain a roughly constant number of neighbours. These larger smoothing
lengths will enhance the density contribution at high z. When this numerical effect is borne in
mind, both simulations demonstrate that material directly thrown above the midplane by the
stream-disk impact is likely to dominate the absorption column at high z.
3.3. Column Densities Towards the Primary
-- 11 --
Fig. 5. -- Density from simulation LMXB2 (full Roche potential, q = 0.2) in the x − y plane at (a)
z = 0, and (b) z = 0.12.
In order to quantify the observable effects of the material thrown out of the plane we
have calculated the column density σ along various lines of sight toward the primary from our
simulations. We calculate,
σ(φ, θ) = Z ρdl
(5)
from integration of the SPH density estimator along lines of sight toward the primary at phase φ
and at angle θ above the disk midplane. φ values are chosen so that neighbouring lines of sight are
calculated from different sets of particles and hence independent. As the innermost particles have
overly large smoothing lengths for the reasons already discussed, these are removed before column
densities are calculated. This cut affects only ∼ 1% of the particles.
Fig. 6 shows the results for both simulations, for angles of elevation θ = 0, 6, 12◦. For
comparison, the dipping source X1822-371 is believed to be viewed at an inclination angle of
82o -- 87o (Mason 1989). For LMXB1 there is no variation evident above the noise at θ = 0◦. By
θ = 6◦ the overflowing stream material, seen in Fig. 4 at high z, is producing a significant
contribution to the column density, leading to a peak in σ at phase φ ∼ 0.8. There is marginal
evidence for a broad feature also at φ ∼ 0.2, caused by the compressed disk rim expanding
outwards downstream of the stream impact point. However, at the highest angle of elevation,
θ = 12◦, the overflowing stream gas provides the only source of absorbing column, and there is
a single strong peak centred on φ = 0.85 and extending from phase 0.7 through to eclipse. For
an opacity generated solely by Thomson scattering a column density of order unity is required to
produce an optical depth of 1, and hence for this accretion rate (3× 10−9 M⊙ yr−1) the overflowing
-- 12 --
Fig. 6. -- Column density σ towards the primary as a function of phase for varying angles of
elevation above the disk plane θ = 0, 6, 12 degrees. Left panels are for simulation LMXB1, right
panels for LMXB2. Units are gcm−2. Error bars are calculated from 4 independent time-slices of
the simulations.
gas would be sufficient to create significant absorption even at angles to the midplane θ >
∼ 10◦.
For run LMXB2 the situation is more complicated. At θ = 12◦ there is a single strong peak
in column density at phase 0.8, which as before is associated with the stream-disk interaction.
There may be a hint of a very small peak at φ ∼ 0.1 -- 0.2. However in this simulation there is
significant variation in σ with phase even at θ = 0◦, as a result of the marked distortion of the
disk shape under the gravitational influence of the companion. Of course viewed directly edge-on,
the optical depth to the central object is enormous, and so the variation in this regime would
not be observable. These variations suggest, however, that at intermediate viewing angles both
effects should be important, and this is seen in the plot for θ = 6◦, where two clear peaks are
produced -- one arising from the stream overflow and one from the disk structure. Since the disk
precesses in the frame of the binary (for low q systems), the absorption feature arising solely
from the eccentric disk structure should vary linearly in phase with time. Conversely, structures
-- 13 --
Fig. 7. -- Destination of stream mass (in arbitrary units) as a function of radius.
generated directly by the stream-disk interaction should remain at fixed phases (regardless of the
mass ratio of the binary).
3.4. Fate of Stream Material
Time-dependent models for accretion disks in (dwarf nova) systems vulnerable to thermal
instability-driven outbursts generally add mass to the disk from the stream either in the outermost
computational zone (e.g. Pringle, Verbunt & Wade 1986), or else with a narrow gaussian profile
at the outer edge (e.g. Cannizzo 1993). Neither is likely to be a good representation of the
physical situation if a significant fraction of the stream mass overflows the disk rim and circularises
at smaller radii. To examine this possibility, Fig. 7 shows the destination of stream material
as a function of radius in the disk, for simulation LMXB1, where the disk is circular and the
complicating effects of an eccentric disk are not present. The distribution is bimodal, with most
of the mass being deposited and circularising at the outer disk rim, but with a significant fraction
(∼ 1/3) ending up at much smaller radii. The outer peak is caused by the dense inner core of the
stream, which on impact with the disk becomes buried in the outer edge, while the inner peak
-- 14 --
(near the stream circularisation radius) is formed from the stream gas that overflows the disk rim.
This result must be treated with some caution however, as the current simulations do not
have sufficient resolution to resolve, for example, any surface instabilities that might entrain the
overflowing stream gas more efficiently and thus keep it at larger radii. Calculations with a higher
resolution and better treatment of the thermal properties of the gas will be required to investigate
this possibility. The general picture from our simulations, however, is that neither mass addition
exclusively at the outer edge nor mass addition primarily at the circularization radius is likely to
prove a good approximation. Rather, the surface density increment is bimodal with roughly equal
amplitudes at the two radii, and we anticipate that this is likely to alter the detailed behaviour
seen in time-dependent disk instability models.
3.5. Location of energy dissipation
In our simulations artificial viscosity is used in the momentum equation to generate shocks
and prevent particles streaming freely through each other. In the physical disk, these shocks would
lead to strong heating of the gas, and create bright spots on the disk surface. The strength of this
dissipation can be estimated by calculating the shock heating that would occur in the simulation
if the gas had an adiabatic (γ = 5/3) equation of state, while the flow pattern remained as we
have described.
Fig. 8 shows the energy dissipation estimated in this way as a function of phase, for annuli
that cover the radial extent of the disk. For the annuli at smaller radii, the background level of
dissipation is uniformly greater because of the increased viscous dissipation that is unconnected
with shocks. Excesses over this level represent areas where bright spots might be expected to
occur. As expected, the strongest dissipation is found in the outermost annulus, at the location
where the stream strikes the disk. However there is also strong shock heating occuring in the
innermost ring, at the location where the overflowing stream gas collides again with the disk (at
phase φ ∼ 0.6). A smaller fraction of stream material reaches this radius, but being being deeper
in the potential well the total shock heating is similar. No prominent features are seen in the two
intermediate annuli.
The width of the features seen in Fig. 8 represents only the azimuthal extent of the regions
where shock heating would occur, the actual size of the bright spots produced as the shocked gas
cools will undoubtably be much larger and may cover a sizeable fraction of the disk circumference
(Rozcyczka & Schwarzenberg-Czerny 1987). Our calculations suggest that a second bright spot
should be formed at the roughly diametrically opposed phase where the overflowing gas returns
to the disk plane. This inner bright spot is of comparable luminosity to that formed directly by
the stream-disk impact, though relative to the background emission at the smaller radius it is,
of course, much weaker. It would thus be most easily seen in systems with low viscosity (e.g.
-- 15 --
Fig. 8. -- Rate of heating from shocks as a function of phase for annuli in the disk (simulation
LMXB1). The panels run from the innermost annulus at upper left, to the outermost at bottom
right. The disk edge is at R ≈ 0.9 in these units.
dwarf novae with long recurrence times), and could provide valuable clues to the degree of stream
overflow occuring in these systems.
4. SUMMARY AND DISCUSSION
We have presented three dimensional SPH calculations of the stream-disk interaction. The
main conclusion to be drawn from these simulations is that significant amounts of stream material
are able to overflow or bounce off the edge of the disk, and flow over the disk surface towards
smaller radii. This creates a peak in the column density towards the primary at a phase of around
0.8, and is a likely cause of the absorption dips observed at around this phase in some nearly
edge-on low-mass X-ray binaries (see e.g. review by White, Nagase, & Parmar 1995). If this is
the case, the absorbing material is located (in radial distance) intermediate between the disk edge
-- 16 --
∼ 10◦), we see no significant absorbing column at any other phases. Recent EUVE observations
and the stream circularization radius. At the highest angles of elevation above the disk surface
(θ >
of U Gem in outburst (Long et al. 1996) show an absorption dip (at phase ∼ 0.75) caused by
material a few scale heights above the disk surface, in complete agreement with the results of the
present study. ASCA observations show that the dips persist in quiescence (Szkody et al. 1996).
This result differs from that of Hirose, Osaki & Minishige (1991), who in addition to predicting
absorption at phase 0.8 also found absorption at phase 0.2 in their models (although as they did
not compute column densities from their simplified calculation it is unclear whether both those
features would be visible at large θ). We do find significant absorption at around phase 0.1 -- 0.2
for lower angles of elevation above the disk plane, and this seems to be primarily due to the
eccentricity induced in the disk by the gravity of the companion. We note that in low mass ratio
systems absorption features generated by the eccentric disk structure should precess with the disk,
and thus can be separated from those induced by the stream-disk interaction, which remain at
a fixed phase (∼ 0.8) regardless of the mass ratio. Therefore, in systems where two absorption
features are seen, a straightforward observational test to determine their origin is possible, by
monitoring for a prolonged period the phases of maximum absorption.
The presence of gas that is not in hydrostatic equilibrium flowing over the disk raises a
number of intriguing possibilities. If exposed to an X-ray power-law spectrum from the central
source, it may be prone to a two-phase instability and break up into dense clouds surrounded
by a hot diffuse medium (Krolik, McKee & Tarter 1981; Frank, King & Lasota 1987). Such a
picture might explain the very short-timescale variations seen in some dipping X-ray sources.
Less speculatively, the wide range of radii over which mass from the stream circularises in our
simulations, implies that significant changes might be expected in the behavior predicted from
time-dependent thermal disk instability models. In particular, during quiescence, the quasi
ballistic flow over the disk might dominate inflow caused by viscosity in the outer disk, and this
would affect predictions for the behaviour of the disk radius. Significant stream overflow implies
that less material with low specific angular momentum will be added at the outer edge of the disk,
and hence less contraction of the disk radius during quiescence would be expected. If the degree of
overflow were large (as might occur if the outer disk were cold relative to the mass-donating star),
then substansive differences in the surface density profile during quiescence would also occur. This
could lead, for example, to thermal instabilities triggered at intermediate disk radii, and other,
detailed, differences to standard models.
For systems that undergo outbursts, the presence of overflowing stream gas provides a possible
observational probe of the disk radius. As the disk expands, the phase of maximum absorption
is expected to increase (closer to the phase of eclipse) as a consequence both of the position and
angle at which the accretion stream reaches the disk. More subtle effects might also occur due
to the varying temperature of the outer disk over the course of an outburst cycle. A hotter disk
is thicker, and this should reduce the amount of stream material able to overflow the rim and
thereby reduce the depth of the absorption. A thicker disk might of course absorb radiation from
-- 17 --
the primary directly, but this effect would be axisymmetric and thus distinguishable from that
caused by high z stream gas. Simultaneously more energy would be dissipated in the bright spot
at the disk rim, so that the bright spot luminosity would increase as the outburst commenced.
There is some observational evidence that this brightening does occur (Livio 1994).
We should note that in their attempt to model the visual light curve of the eclipsing supersoft
X-ray source CAL 87, Schandl, Meyer-Hofmeister, & Meyer (1996), found that they had to assume
the presence of an optically thick "spray" of material around the disk. They conjectured that
such a spray might form by the disk-stream interaction. An examination of Fig. 2 reveals that we
indeed obtain a configuration similar to that assumed by Schandl et al. (1996).
Finally we remark that in this paper we have concentrated on explaining features seen in the
X-ray and EUV wavebands. Important information is also available in the optical, especially from
the analysis of time-resolved spectra using doppler tomography (Marsh & Horne 1988). Many
systems have now been mapped using this technique (e.g. Kaitchuck et al. 1994), and in some
cases the path of the accretion stream can be clearly traced. A future comparison of synthetic
doppler maps generated from simulations of the type presented here with these data appear to
provide a fruitful direction in which the research into stream-disk interactions can continue.
ACKNOWLEDGEMENTS
We thank Ian Bonnell, Cathie Clarke, Melvyn Davies, Andy Fabian, Steve Lubow and Donald
Lynden-Bell for useful discussions and suggestions. PJA thanks Space Telescope Science Institute
for hospitality. ML acknowledges support from NASA Grant NAGW-2678 at ST ScI, and thanks
the Institute of Astronomy, Cambridge, for its hospitality.
-- 18 --
REFERENCES
Bate, M. 1995, Ph.D. Thesis, University of Cambridge, p. 43
Bate, M. 1996, in preparation
Benz, W. 1990, in The Numerical Modeling of Nonlinear Stellar Pulsations, ed. J.R. Buchler,
Kluwer Academic Publishers, Dordrecht, p. 269
Cannizzo, J. K. 1993, ApJ, 419, 318
Davies, M.B., Ruffert, M., Benz W., & Muller, E. A&A, 272, 430
Dgani, R., Livio, M., & Soker, N. 1989, ApJ, 336, 350
Frank, J., King, A. R., & Lasota, J. P. 1987, A&A, 178, 137
Hellier, C., & Livio, M. 1994, ApJ, 424, L57
Hirose, M., Osaki, Y., & Minishige, S. 1991, PASJ, 43, 809
Kaitchuck, R. H., Schlegel, E. M., Honeycutt, R. K., Horne, K., Marsh, T. M., White, J. C., II, &
Mansperger, C. S. 1994, ApJS, 93, 519
Krolik, J. H., McKee, C. F., & Tarter, C. B. 1981, ApJ, 249, 422
Lanzafame, G., Belvedere, G., & Molteni, D. 1994a, MNRAS, 258, 152
Lanzafame, G., Belvedere, G., & Molteni, D. 1994b, MNRAS, 267, 312
Lin, D. N. C., & Papaloizou, J. 1979, MNRAS, 186, 799
Lin, D. N. C., & Pringle, J. E. 1976, in Structure and Evolution of Close Binary Systems, ed. P.
Eggleton, D Reidel, Dordrecht, p. 237
Livio, M. 1993, in Accretion Disks in Compact Stellar Systems, ed. J. C. Wheeler (Singapore:
World Scientific), p. 243
Livio, M. 1994, in Interacting Binaries, Saas-Fee Advanced Course 22, eds. H. Nussbaumer & A.
Orr, Springer-Verlag, Berlin Heidelberg, p. 142
Livio, M., Soker, N., & Dgani, R. 1986, ApJ, 305, 267
Long, K. S., Mauche, C. W., Raymond, J. C., Szkody, P., & Mattei, J. A. 1996, ApJ, in press
Lubow, S. H. 1989, ApJ, 340, 1064
Lubow, S. H. 1991, ApJ, 381, 268
Lubow, S. H., & Shu, F. H. 1975, ApJ, 198, 383
Lubow, S. H., & Shu, F. H. 1976, ApJ, 207, L53
Marsh, T. R., Horne, K. 1988, MNRAS, 235, 269
Mason, K. O. 1989, in Proc. 23rd ESLAB Symp. on Two-Topics in X-Ray Astronomy, ESA
SP-296, p. 113
-- 19 --
Monaghan, J. J. 1992, ARA&A, 30, 543
Murray, J. R. 1996, MNRAS in press
Paczy´nski, B. 1977, ApJ, 216, 822
Papaloizou, J., & Pringle, J. E., 1977, MNRAS, 181, 441
Pringle, J. E. 1981, ARA&A, 19, 137
Pringle, J. E., Verbunt, F., & Wade, R. A. 1986, MNRAS, 221, 169
Rozcyczka, M. & Schwarzenberg-Czerny, A. 1987, Acta Astron., 37, 141
Schandl, S., Meyer-Hofmeister, E., & Meyer, F. 1996, A&A, submitted
Shakura, N.I., & Sunyaev, R.A. 1973, A&A, 24, 337
Szkody, P., Long, K. S., Sion, E. M., & Raymond, J. C. 1996, ApJ, in press
Warren, J. K., Sirk, M. M., & Vallerga, J. V. 1995, ApJ, 445, 909
White, N. E. & Holt, S. S. 1982, ApJ, 257, 318
White, N. E., Arnaud, K., Day, C. S. R., Ebisawa, K., Gotthelf, E. V., Mukai, K., Soong, Y.,
Yaqoob, T., & Antunes, A. 1994, PASJ, 46, L97
White, N. E., Nagase, F., & Parmar, A. N. 1995, in X-Ray Binaries, eds. W. H. G. Lewin, J. van
Paralijs, & E. P. J. vanden Heuvel (Cambridge: Cambridge University Press), p. 1
Whitehurst, R. 1988a, MNRAS, 232, 35
Whitehurst, R. 1988b, MNRAS, 233, 529
This preprint was prepared with the AAS LATEX macros v3.0.
|
astro-ph/0412326 | 1 | 0412 | 2004-12-14T11:39:44 | Occurence and Luminosity Functions of Giant Radio Halos from Magneto-Turbulent Model | [
"astro-ph"
] | We calculate the probability to form giant radio halos (~ 1 Mpc size) as a function of the mass of the host clusters by using a Statistical Magneto-Turbulent Model (Cassano & Brunetti, these proceedings). We show that the expectations of this model are in good agreement with the observations for viable values of the parameters. In particular, the abrupt increase of the probability to find radio halos in the more massive galaxy clusters (M > 2x10^{15} solar masses) can be well reproduced. We calculate the evolution with redshift of such a probability and find that giant radio halos can be powered by particle acceleration due to MHD turbulence up to z~0.5 in a LCDM cosmology. Finally, we calculate the expected Luminosity Functions of radio halos (RHLFs). At variance with previous studies, the shape of our RHLFs is characterized by the presence of a cut-off at low synchrotron powers which reflects the inefficiency of particle acceleration in the case of less massive galaxy clusters. | astro-ph | astro-ph | Journal of The Korean Astronomical Society
37: 1 ∼ 5, 2004
Occurence and Luminosity Functions of Giant Radio Halos from Magneto-Turbulent
Model
R. Cassano1
,
1 Dipartimento di Astronomia ,Univ. di Bologna, Italy
2 Istituto di Radioastronomia del CNR ,Bologna, Italy
2, G. Brunetti2, G. Setti1
2
,
4
0
0
2
c
e
D
4
1
1
v
6
2
3
2
1
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
We calculate the probability to form giant radio halos (∼1 Mpc size) as a function of the mass of the host
clusters by using a Statistical Magneto-Turbulent Model (Cassano & Brunetti, these proceedings). We show that
the expectations of this model are in good agreement with the observations for viable values of the parameters. In
particular, the abrupt increase of the probability to find radio halos in the more massive galaxy clusters (M >∼2 ×
1015M⊙) can be well reproduced. We calculate the evolution with redshift of such a probability and find that giant
radio halos can be powered by particle acceleration due to MHD turbulence up to z∼0.5 in a ΛCDM cosmology.
Finally, we calculate the expected Luminosity Functions of radio halos (RHLFs). At variance with previous studies,
the shape of our RHLFs is characterized by the presence of a cut-off at low synchrotron powers which reflects the
inefficiency of particle acceleration in the case of less massive galaxy clusters.
Key Words : acceleration of particles - turbulence - radiation mechanism: non-thermal, galaxy clusters: general
-radio continuum
I.
Introduction
Radio observations of galaxy clusters indicate that
the detection rate of radio halos (RHs) shows an abrupt
increase with increasing the X-ray luminosity of the
host clusters. In particular about 30-35% of the galaxy
clusters with X-ray luminosity larger than 1045 erg/s
show diffuse non-thermal radio emission (Giovannini
& Feretti 2002); these clusters have also high temper-
ature (kT > 7 keV) and large mass (>∼ 2× 1015M⊙).
Furthermore giant RHs are frequently found in merging
clusters (e.g., Schuecker et al 2001). These observations
suggest that there is a connection between thermal and
non-thermal phenomena in galaxy clusters.
Recent papers (Ensslin and Rottgering 2003; Kuo et
al. 2004) have investigated the statistics of RHs and
their connection with the thermal properties of the host
clusters from a theoretical point of view. These works
are based on assumpions in defining the condition of
RHs formation from observational correlations and/or
mass thresholds. Present data suggest that giant RHs
may be accounted for by synchrotron emission from
relativistic electrons reaccelerated by the turbulence
generated in the cluster volume during merger events
(Brunetti 2003; Brunetti this proceedings). Thus, with
the aim to investigate the statistical properties and the
connection between thermal and non-thermal phenom-
ena in galaxy clusters, we have developed a statistical
magneto-turbulent model (Cassano & Brunetti 2004,
C&B model; Cassano & Brunetti these proceedings)
in which we follow the formation of clusters of galaxies
(making use of the extended Press & Schechter (1974)
formalism) and estimate the injection of fluid turbu-
lence and of fast magnetosonic (MS) waves during clus-
ter mergers. Then we calculate the evolution of the
electron spectra in the ICM and the resulting radio
(synchrotron) and hard X-ray (Inverse Compton) emis-
sion spectra. By using our model we can thus investi-
gate the probability of formation of RHs in a well de-
fined physical framework, the evolution with redshift of
such a probability and the expected Luminosity Func-
tions of RHs (RHLFs). Here we applay the C&B model
under the following assumptions:
- We focus the attention on the formation of giant
RHs only (radius RH ∼ 500 h−1
50 Kpc).
- The magnetic field strength averaged over the
emitting volume is assumed to be < B >∼ 0.5
µG, independent on the mass of the parent clus-
ter.
The adopted cosmologies are: EdS (Ho = 50 Km
s−1M pc−1, Ωo,m = 1) and ΛCDM (Ho = 70 Km
s−1M pc−1, Ωo,m = 0.3, ΩΛ = 0.7, σ8 = 0.9).
II. Occurence of RHs: predictions vs observa-
tions
By making use of the C&B model we have run Monte
Carlo simulations to obtain a sufficiently large number
of merger trees in order to have a large synthetic pop-
ulation of galaxy clusters with a wide range of present
day masses and temperatures.
In this way we are
able to statistically follow the cosmological evolution
of the non-thermal emission and of the properties of
the thermal ICM. Clusters with RHs in our synthetic
population are identified with those objects with a syn-
chrotron cut-off νb>∼102 MHz in a region of 1 M pc h−1
50
size. We have calculated the probability to form RHs
(z ≤ 0.2) in two mass bins: binA=[1.8 − 3.6] × 1015M⊙
and binB=[0.9−1.8]×1015M⊙ (EdS cosmology). These
mass bins are consistent with those considerated in the
-- 1 --
2
R.Cassano, G.Brunetti, G. Setti
Fig.
1. -- Expected formation probability of RHs
(RH ≃ 500h−1
50 kpc, B ∼ 0.5µG) as a function of param-
eter ηt in a EdS cosmology (solid lines with error bars)
and in a ΛCDM cosmology (dotted lines) in the mass
bins: binA=[1.8 − 3.6] 1015M⊙ h−1
50 and binB=[0.9 −
1.8] 1015M⊙ h−1
50 for EdS case and binA=[1.9−3.8]·1015
M⊙h−1
70 for the
ΛCDM model. The two bottom dashed lines mark the
observed probabilities for RHs in the mass binB while
the two top dashed lines mark the observed probabil-
ities in the mass binA; observational regions already
account for 1σ errors.
70 and binB=[0.945 − 1.9] · 1015 M⊙h−1
observational studies and thus allow us to compare our
expectations with observations.
In Fig.(1) we report
the probability to form a giant (≃1 Mpc h−1
50 size) RH
(red points) in the two mass-bins (including the statisti-
cal error estimated from our Montecarlo simulations) as
a function of the parameter ηt, which gives the fraction
of energy of the turbulent motions injected by cluster
merger which is channeled in the form of MS waves in
the C&B model.
The dashed blue lines mark the range of the observed
probabilities (Giovannini et al. 1999) in the binA (top
dashed region) and in the binB (bottom dashed region),
respectively. For a comparison in Fig.(1) we also report
the probability to form a RH (z ≤ 0.2) in a ΛCDM
cosmology (green dotted lines). As expected, we find
that at z ≤ 0.2 the results are relatively indipendent
from the considered cosmology, with the ΛCDM model
being slightly less efficient. The main result is that
in both EdS and ΛCDM models it is possible to find
a unique interval of ηt in which the model reproduces
the observed probability for both the cluster-mass bins.
Fig. 2. -- Probability to form giant RHs (> 1M pc h−1
50
size) as a function of the cluster mass in two relevant
redshift bins: z=0-0.2 (red lines) and z=0.2-0.4 (black
lines) in a ΛCDM model. Vertical dashed lines mark
the [1.9 − 3.8] × 1015M⊙ mass bin of Fig.(1). Calcula-
tions are obtained for ηt = 0.3.
In particular, in agreement with observations and in-
dipendently from the adopted cosmology, we find that
20-30% of clusters in the binA can form a RH and that
only 2-3% of galaxy clusters in the binB host a RH.
Given the requested values of ηt (Fig.(1)), we find that
the relatively high occurence of RHs observed in mas-
sive clusters can be well reproduced by our particle ac-
celeration model under reasonable conditions, i.e. that
a fraction of 20-30% of the energy of the turbulent mo-
tions injected during cluster merger (which corresponds
to a few percent of the thermal energy) is in the form
of MS waves.
III. Radio Halos Statistics and cluster mass
In the previous Section we have found that the
probabilty to form a RH has a strong dependence on
the mass of the host cluster, it goes from few % for
M < 1.9 × 1015M⊙ to 20-30% for M >∼2 × 1015M⊙, in
agreement with observations (Fig.(1)).
Now we calculate the probability to find RHs as a
function of the cluster mass from our synthetic popu-
lation of galaxy clusters in a ΛCDM cosmology: this
is a crucial point of our model and marks the differ-
ence with previous studies (e.g., Ensslin & Rottgering
(2002)). As an example, in Fig.(2) we report this prob-
ability in two redshift bins z=0-0.2 (black lines) and
z=0.2-0.4 (red lines). Clusters with mass << 1015M⊙
Occurence and Luminosity Functions of Giant Radio Halos from Magneto-Turbulent Model 3
mergers between massive subclusters are expected to
inject turbulence on larger volumes, of the order of VH ,
and thus the efficiency of the generation of RHs is not
reduced and this further favour massive objects as the
parent clusters of RHs.
More quantitatively,
it can be shown (Cassano
& Brunetti 2004) that the acceleration efficiency χ
(within VH ), triggered by a major merger event scales
about with χ ∝ M 0.75−1.25 (0.75 for M ≥ 3 · 1015M⊙,
1.25 for M < 1015M⊙). Since the maximum energy
of the accelerated electrons is γb ∝ χ/(B2 + B2
CM B),
where BCM B = 3.2 · (1 + z)4µG is the strength of the
equivalent magnetic field of the CMB, and the break
frequency is νb ∝ γ 2
b B, one has:
Fig. 3. -- Probability to form giant RHs (∼ 1M pc
h−1
50 size) with redshift in the mass bin [1.9 − 3.8] ×
1015M⊙h−1
70 in a ΛCDM cosmology. Calculations are
performed for ηt = 0.3.
have a negligible probability to form giant RHs; on the
other hand, such a probability is found to reach ∼40%
for clusters with masses > 3 × 1015M⊙ in the 0-0.2 red-
shift bin. For a comparison with Fig.(1) we also mark
in Fig.(2) the mass bin [1.9 − 3.8] × 1015M⊙ (vertical
blue dashed lines).
Thus, the important finding of these calculations is
that only massive clusters can host giant RHs (RH ≥
500 kpc h−1
50 ) and that the probabilty to form these dif-
fuse radio sources presents an abrupt increase for clus-
ters with about M >∼2 × 1015M⊙ (Fig.(1) and Fig.(2)).
These findings can be simply explained in the frame-
work of our model.
it can be shown that
in the C&B model the energy of the turbulence in-
jected in galaxy clusters is expected to roughly scale
with the thermal energy of the clusters (Cassano &
Brunetti 2004; Cassano & Brunetti these proceeding
Fig(2)). This seems reasonable and immediately im-
plies that the energy density of the turbulence is an
increasing function of the mass of the clusters,
i.e.
Et ∝ T ∝ M a (a=0.56-0.67), and thus particle acceler-
ation is favoured in massive clusters.
Infact,
In general, the infall of subclusters through a main
cluster, which is not very massive, injects turbulence in
a volume Vt (calculated using Ram Pressure Stripping,
Fujita, Takizawa, Sarazin 2003; Cassano & Brunetti
2004) which is found to be smaller than that of giant
RHs, VH (VH = 4πR3
H /3), and thus the efficiency of the
mechanism is reduced by about a factor of Vt/VH in the
case of less massive clusters. On the other hand, major
νb ∝ M 1.5−2.5
B
(B2 + B2
CM B)2
(1)
and consequently massive clusters are statistically
favourite to have νb>∼102 MHz (which is the adopted
condition to define the presence of a RH).
IV. Evolution of radio halos with redshift
The probability to form RHs dependes on the combi-
nation of the energy losses suffered by relativistic elec-
trons (mainly due to IC losses ∝ (1 + z)4) with the
acceleration efficiency powered by the turbulence gen-
erated during cluster mergers (which depends on the
merger history).
In this Section we calculate the evolution with red-
shift of the probability to form RHs in a ΛCDM cosmol-
ogy. As an example, in Fig.(3) we report the probabil-
ity to form a giant RH (≃ 700 Kpc in a ΛCDM cosmol-
ogy) with redshift in the mass bin: [1.9 − 3.8] × 1015M⊙
for a representative value of ηt (ηt = 0.3). The occur-
rence of RHs decreases with redshift due to the higher
IC energy losses. We note however that such a de-
crease is not dramatic since in a ΛCDM Universe ma-
jor mergers develope at slightly higher redshift with
respect to a EdS Universe. For instance, in the con-
sidered case the formation rate of RHs is 20-36% at
relatively low redshift and decreases to 10% at higher
redhifts (0.3 ≤ z ≤ 0.4).
V. The Luminosity Functions of Radio Halos
(RHLFs)
We have alredy shown that the observed probability
to find RHs with the cluster mass is well reproduced
by the C&B model (see Sec.2). In Cassano & Brunetti
2004 (see also Cassano & Brunetti these proceedings)
it has also been shown that the typical synchrotron
and IC luminosity of RHs can be well reproduced by
the model assuming that during mergers a few percent
of the thermal energy of the cluster is in the form of
4
R.Cassano, G.Brunetti, G. Setti
RHs per Gpc3 as a function of the radio power) ex-
pected by our model and the expected RHLFs in the
bins 0 ≤ z ≤ 0.2 and 0.2 ≤ z ≤ 0.4 (lines with points).
Our RHLFs are compared with the Local (RHLF )E&R
(blue solid line) of Ensslin & Rottgering (2002). The
(RHLF )E&R is obtained combining the X-ray observed
luminosity function of clusters with the radio luminos-
ity - X-ray luminosity correlation and assuming that
a costant fraction frh = 1/3 of galaxy clusters have
RHs. The most important difference between the two
luminosity functions is that our RHLF shows a cut-
off/flattening at low radio powers. We stress that the
flattening at low powers is a unique feature of particle
acceleration models since it marks the effect of the de-
crease of the efficiency of the particles acceleration in
the case of the less massive galaxy clusters and conse-
quently the presence of a synchrotron cut-off νb < 102
MHz.
Future radio observations (e.g., with LOFAR and
LWA) should be able to test the presence of such a
low-power cut-off in the RHLFs and the evolution of
the RHLFs with redshift.
ACKNOWLEDGEMENTS
G.B. and R.C. acknowledge partial support from
CNR grant CNRG00CF0A.
REFERENCES
Arnaud M., Evrard A.E., 1999, MNRAS 305, 631.
Brunetti G., 2003, in 'Matter and Energy in Clusters of
Galaxis', ASP Conf. Series, vol.301, p.349, eds. S.
Bowyer and C.-Y. Hwang
Cassano R., Brunetti G., 2004 submitted to MNRAS.
Ensslin T. A., Rottgering H., 2002, A&A, 396, 83.
Fujita Y., Takizawa M., Sarazin C.L., 2003, ApJ 584, 190
Giovannini G., Tordi M., Feretti L., 1999, NewA 4, 141.
Giovannini G., Feretti L., 2002 in 'Merging Processes
L.Feretti,
in Galaxy Cluster', vol.272, p.197, eds.
I.M.Gioia, G.Giovannini.
Feretti L., 2003,
in 'Matter and Energy in Clusters of
Galaxis', ASP Conf. Series, vol.301, p.143, eds. S.
Bowyer and C.-Y. Hwang.
Kuo P.-H., Hwang C.-Y., Ip W.-H., 2004 ApJ 604, 108.
Fig. 4. -- RHLFs (nH (P ) × P ) expected by the C&B
model. Results are shown for the redshift bins 0 ≤ z ≤
0.2 (black curve with points) and 0.2 ≤ z ≤ 0.4 (red
curve with points). The expected Local RHLF (blue
curve with points) is also reported together with the
Local RHLF from Ensslin & Rottgering (2002) (solid
blue line) for a comparison.
MS waves (i.e., ηt > 0.1 − 0.2). Given these promis-
ing results, in this Section we derive the expected lu-
minosity functions of giant RHs (RHLFs). First we
use the probility to form RHs with the cluster's mass
P ∆z
∆M (Fig.2) to estimate the mass functions of RHs
(dNH (z)/dM dV ):
dNH (z)
dM dV
=
dNcl(z)
dM dV
× P ∆z
∆M = nP S × P ∆z
∆M ,
(2)
where nP S = nP S(M, z) is the Press & Schechter
(1974) mass function (we use nP S since our model is
based on Press & Schechter formalism). The RHLF is
given by:
dNH(z)
dV dP1.4
=
dNH (z)
dM dV (cid:30) dP1.4
dM
.
(3)
Press W.H., Schechter P., 1974, ApJ 187, 425.
Schuecker P., Bhringer H., Reiprich, T.H., Feretti L., 2001,
A&A 378, 408.
In order to derive dP1.4/dM in Eq.(3), we combine
the observed correlations between radio power at 1.4
Ghz (P1.4GHz ) and bolometric X-ray luminosity (LX )
(e.g., Feretti 2003) and between LX and the virial mass,
M200 (e.g., Arnaud & Evrard 1999). The used P1.4GHz-
M correlation is obtained collecting the data from all
the known clusters with giant RHs and converting them
in a ΛCDM cosmology. In Fig.(4) we report the Lo-
cal (here calculated for z < 0.05) RHLF (number of
|
astro-ph/9902300 | 3 | 9902 | 1999-03-27T22:43:01 | EUV images of the Clusters of Galaxies A2199 and A1795: clear evidence for a separate and luminous emission component | [
"astro-ph"
] | Since all of the five clusters of galaxies observed by the Extreme Ultraviolet Explorer deep survey telescope are found to possess a diffuse EUV emitting component which is unrelated to the virial gas at X-ray temperatures, the question concerning the nature of this new component has been a subject of controversy. Here we present results of an EUV and soft X-ray spatial analysis of the rich clusters Abell 2199 and 1795. The EUV emission does not resemble the X-ray morphology of clusters: at the cluster core the EUV contours are organized, but at larger radii they are anisotropic, and therefore need not have originated from a hydrostatically stable medium. The ratio of EUV to soft X-ray intensity rises with respect to cluster radius, to reach values ~10 times higher than that expected from the virial gas. The strong EUV excess which exists in the absence of soft X-ray excess poses formidable problems to the non-thermal (inverse-Compton) scenario, but may readily be explained as due to emission lines present only in the EUV range. In particular, warm gas produced by shock heating could account for such lines without proliferation of bolometric luminosities and mass budgets. | astro-ph | astro-ph | EUV images of the Clusters of Galaxies A2199 and A1795: clear
evidence for a separate and luminous emission component
Richard Lieu 1, Massimiliano Bonamente 1, Jonathan P. D. Mittaz 2
1Department of Physics, University of Alabama, Huntsville, AL 35899, U.S.A.
2Mullard Space Science Laboratory, UCL, Holmbury St. Mary, Dorking, Surrey, RH5 6NT,
U.K.
Received
;
accepted
9
9
9
1
r
a
M
7
2
3
v
0
0
3
2
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
-- 2 --
ABSTRACT
Since all of the five clusters of galaxies observed by the Extreme Ultraviolet
Explorer (EUVE) deep survey telescope are found to possess a diffuse EUV
emitting component which is unrelated to the hot intracluster medium (ICM) at
X-ray temperatures, the question concerning the nature of this new component
has been a subject of controversy. Here we present results of an EUV and
soft X-ray spatial analysis of the rich clusters Abell 2199 and 1795. The EUV
emission does not resemble the X-ray morphology of clusters: at the cluster core
the EUV contours are organized; at larger radii they are anisotropic, and are
therefore unrelated to the hot ICM. The ratio of EUV to soft X-ray intensity
rises with respect to cluster radius, to reach values ∼ 10 times higher than
that expected from the hot ICM. The strong EUV excess which exists in the
absence of soft X-ray excess poses formidable problems to the non-thermal
(inverse-Compton) scenario, but may readily be explained as due to emission
lines present only in the EUV range. In particular, warm gas produced by
shock heating could account for such lines without proliferation of bolometric
luminosities and mass budgets.
The 'cluster soft excess' (CSE) phenomenon, which originated from EUVE, was
confirmed by the ROSAT and BeppoSAX (Lieu et al 1996a,b; Bowyer, Lampton and
Lieu 1996; Fabian 1996; Mittaz, Lieu and Lockman 1998; Bowyer, Lieu and Mittaz 1998;
Kaastra 1998). The non-thermal interpretation of this phenomenon (Ensslin and Biermann
1998; Hwang 1997; Sarazin and Lieu 1998), favored on plausibility grounds over the original
thermal (warm gas) scenario, is recently supported by the BeppoSAX and RXTE discoveries
of hard X-ray tails in the spectra of several clusters (Kaastra 1998; Fusco-Femiano et al
1998; Rephaeli et al 1999), some of which exhibit similar radial trends as the CSE. In
-- 3 --
this Letter we report first results on a spatial analysis of the EUV which reveals clear and
essential differences in spatial morphology between the EUV and soft X-rays. Our results
concern the rich clusters Abell 2199 and 1795; although a similar conclusion exists also in
the case of Virgo, A4038, and Coma, the analyses of these will be presented in a subsequent
and more detailed paper.
Abell 2199 was observed by the EUVE deep survey (DS) Lex/B (69 - 190 eV) filter
for ∼ 50, 000 sec. Figure 1(a) is a radial profile of the cluster surface brightness after
removal of point source contributions and subtraction of a background taken from the > 15
arcmin region where there is no further evidence of cluster emission. Standard correction
procedures were applied1 to recover the signals lost within a small region near boresight of
∼ 1 arcmin radius and ∼ 3 arcmin away from the cluster center - a region known as the
deadspot, which was caused by the February 1993 observation of the intense EUV source
HZ 43. This procedure ensures full integrity in the data reduction procedure; its actual
effect is minor, resulting only in a slight, statistically insignificant enhancement of the
brightness in the central portion of the profile.
To compare with the X-ray behavior, data from a ROSAT Position Sensitive
Proportional Counter pointing which took place in July 1990 were extracted from the
archive2. Again point sources were removed, and a background as determined from the
outermost annulus was subtracted after vignetting corrections. A radial profile of the
EUV to soft X-ray ratio is shown in Figure 1(b). To perform a quantitative assessment
of the trend, we also show in Figure 1(b) the expected value of the ratio if the emission
1See
'Deep
Survey Dead
Spot Correction Algorithm',
internal memo
MMS/EUVE/0084/94, Center for EUV Astrophysics, Berkeley
2HEASARC, available at http://heasarc.gsfc.nasa.gov/docs/rosat/archive.html, and
maintained by the ROSAT/ASCA Guest Observer Facility.
-- 4 --
originated from the virial gas alone. This expected value is obtained by modeling the PSPC
0.2 - 2.0 keV (PH channels 18 - 200) data of concentric annuli with a single temperature
MEKAL thin plasma code (Mewe, Gronenschild & van den Oord 1985; Mewe, Lemen &
van den Oord 1986; Kaastra 1992), using an abundance of 0.5 solar (David et al 1993)
and line-of-sight column density as measured by a dedicated observation at Green Bank
(for further details see Lieu et al 1996a,b), which reported NH = 8.3 × 1019 cm−2 with
a nominal error of < 1019 cm−2. No CSE was evident in the PSPC data: specifically the
model satisfactorily accounts for all the data of the employed PH channels.
However, the Lex/B data indicate substantial CSE, as it can be seen from Figure 1b
that the relative strength of the Lex/B CSE rises with radius, analogous to Abell 1795
(Mittaz, Lieu & Lockman 1998). This is borne-out even more by the 2-D images below. We
emphasize that the effect of concern is only slightly modified if more accurate abundances
(Mushotzky et al 1996) were used to compute the expected softness ratio - the basic notion
of a rising trend in this ratio does not appear to be an abundance issue. Given that most of
the extragalactic EUV radiation is absorbed, and that the DS data are background limited,
the radially rising softness ratio suggests a possible spatial extent of the EUV emission far
larger than our current measurements.
To investigate the spatial distribution of the EUV, we adaptively smoothed the DS
event image with a gaussian filter which encloses a S/N (signal-to-noise ratio3) of ≥ 5 σ. In
the central region of the cluster where the signals are strong, a minimum filter size equal
to that of the DS point spread function (PSF, Sirk et al 1997) was enforced. Contours of
the smoothed and background subtracted surface brightness are shown in Figure 2(a). The
3Since EUVE data are background limited, this S/N is calculated as the ratio (S−B)/√S,
where S is the number of counts enclosed by the filter and B is the detector background
estimated from regions free of cluster emission.
-- 5 --
lowest contour level plotted is ∼ 7.5 % above background; since no part of the entire 2o
× 0.5o area of the Lex/B filter has such a high brightness, other than the cluster region
as shown, the detected emission is not an apparent effect caused by random background
variations.
To further establish the reality of the extended component we produced similarly
smoothed images of (a) two blank field pointings; and (b) an observation of the globular
cluster X-ray source M15 (for 1 -- 6.5 keV detection see Callanan et al 1987), which showed
that the source was visible in the EUV. The uniformity of the background in (a), especially
the absence of any enhancements over a large area centered at boresight - an area where
the observed cluster targets normally occupy, excludes the possibility that diffuse cluster
EUV is a systematic detector spatial effect. In (b), the lack of any extended halo around
M15, despite the source having a peak surface brightness comparable to that of the two
clusters, also rules out leakage of central source radiation in the EUV or X-rays as origin of
the cluster syndrome - a conclusion fully consistent with our understanding of the DS PSF.
All the image data referred to in this Letter are available as FITS files4.
For comparison with spatial behavior in the X-rays, we show in Figure 2(b) a
background subtracted and exposure corrected soft X-ray contour map of A2199, using the
R2 band data of a ROSAT PSPC observation (R2 refers to PH channel 18-41, or ∼ 0.2 -- 0.3
keV by channel boundaries) and a gaussian smoothing filter commensurate in size with the
PSPC PSF in this passband (note that S/N is not a problem for the PSPC data). Despite
the proximity of the R2 and Lex/B passbands, resemblance between Figures 2(a) and 2(b)
is only restricted to the cluster core, where both images show a correlating set of organized
contours. At larger radii the R2 contours remain regular and symmetrical, reminiscent of
4These
can
be
downloaded
via
anonymous
ftp
from the
address
cspar.uah.edu/input/max/.
-- 6 --
emission from a hydrostatic gas, but the Lex/B contours do not share this property.
To assess the statistical significance of the Lex/B anisotropies, we divided the image
into concentric annuli moving outwards from the cluster emission centroid. The surface
brightness of each annulus is computed over octants, and a χ2 test is applied to the
hypothesis that the octant counts are all consistent with their azimuthally averaged
value. In Table 1 we show χ2
red for four equally spaced annuli between 0 and 12 arcmin
radius. It can be seen that for regions outside 6 arcmin radius emission anisotropy is
asserted with a confidence level of > 98.5 % (corresponding to χ2
red ≥ 2.4 for 8 degrees of
freedom). We repeated this test on blank field data of comparable exposure, using the same
near-boresight area of the DS detector where cluster observations were made. Results, also
shown in Table 1, indicate that the detector background distribution does not have similar
anomalies. Furthermore, detailed plots (not shown) reveal that the A2199 anisotropy is due
to deviations of the data from the average value in directions where spatially asymmetric
features are apparent in Figure 2(a). Visual comparison with the optical map of the Digital
Sky Survey revealed no obvious anisotropy in the density and brightness distribution of
member galaxies within the relevant region.
From the radial trend of the Lex/B : R2 ratio (Figure 1b) it was evident that the
relationship between the EUV and the hot ICM decrease with radius. This is borne-out
more strongly in 2-D by Figure 2(c), which indicates an even sharper radial increase of the
ratio in directions normal to the Lex/B brightness contours. The lack of any resemblance
between Figures 2(a) and 2(b) at the outer regions of the cluster provides crucial further
confirmation that the EUV emitting phase leads an existence entirely separate from the hot
ICM. In particular the possibility of interpreting the CSE as due to our overestimate of the
line-of-sight Galactic absorption of the hot ICM radiation (e.g. Arabadjis & Bregman 1998)
is completely excluded by these latest data.
-- 7 --
Apart from A2199, in this Letter we also report results of a similar study of the rich
cluster Abell 1795. An earlier paper (Mittaz, Lieu & Lockman 1998) showed radial profiles
of both the EUV surface brightness and the emission softness ratio for this cluster, it also
showed a feeble CSE in the PSPC data - such information will not be repeated. Here
we focus on a 2-D image analysis, where all the foregoing discussions regarding adopted
techniques apply. As before, we now show in Figure 3 contours of the DS Lex/B and
PSPC R2-band surface brightness, and of the Lex/B : R2 ratio. The maps lead us to
draw conclusions about A1795 similar to those of A2199, viz. that the Lex/B contours are
organized (but differently shaped from the R2 ones) at the core, are anisotropic at larger
radii (see also Table 1) with no similar effects in the optical, and that the rising radial trend
of softness ratio is in agreement with that reported in our earlier paper.
Several interesting physical deductions can be made from a spatial comparison between
the EUV and X-rays. The anisotropic EUV emission could, within the context of the
inverse-Compton scenario, reflect the distribution of the intracluster magnetic field strength
(Ensslin, Lieu & Biermann 1999) in that electrons in regions with > a few µG fields would
have been removed by synchrotron losses as diffusive replenishment is always negligibly slow
(Volk, Aharonian & Breitschwerdt 1996). On the other hand, if the emission is thermal,
as originally advocated, this would imply the existence of warm gas which, by virtue of its
lack of hydrostatic equilibrium at these lower temperatures, could have an anisotropic (and
possibly clumped) distribution as well.
Why do the Lex/B and R2 maps appear so different, given that their energy passbands
have substantial overlap ? Our simulations of the instrumental performances indicate that
this could happen only if the EUV component has a spectrum which does not exceed ∼
0.2 keV. For a non-thermal origin the behavior implies that the bulk of the relativistic
electrons have energies below 200 MeV, a cut-off effect which is most obviously understood
-- 8 --
as due to aging. However, a major difficulty concerns the strong Lex/B excess accompanied
by little or no R2 excess, as this requires an age of ≥ 3 Gyr and hence a very significant
spectral evolution (i.e. losses). By the time the spectrum at the present epoch explains the
observations, at injection the electron pressure alone would have far exceeded that of the
hot ICM gas, leading to an unreasonably short lifetime estimate for the gas.
It is more natural to interpret the marked contrast in the behavior of the Lex/B and
R2 bands as signatures of bright emission lines which are present only in the EUV range.
Although exotic line species arising from an interaction process which may involve dark
matter cannot totally be excluded, a specific scenario of this kind is not available. However,
lines could simply be a direct witness to the presence of thermal gas. For example, warm
under-ionized gas shocked in mixing layers around cool clouds (Fabian 1997) can produce
a blend of intense lines exclusively within the Lex/B energy range, thereby limiting the
bolometric luminosity and (hence) the gas mass budget and mass cooling rate to more
reasonable values. The rise in relative EUV excess with radius may then be a density
scaling effect - at larger radii the rarer gas stays longer in this out of equilibrium phase.
Detailed hydrodynamic simulation of the thermal model, with associated mass implications,
is work in progress.
We thank an anonymous referee for helpful comments.
-- 9 --
Annulus (arcmin) A2199 A1795 Blank field
0-3
3-6
6-9
9-12
1.25
1.53
1.6
2.4
2.62
2.2
2.5
4.7
0.64
0.13
0.4
1.1
Table 1: Reduced χ2 goodness-of-fit test of the azimuthal symmetry in the Lex/B surface
brightness distribution around several concentric annuli. Each annulus was divided into
8 equal octants and the χ2
red value was computed for a model of of constant brightness
given by averaging over all angles. The center of these annuli is either coincident with the
EUV centroid of the relevant cluster or, in the case of the background observation, with a
near-boresight position of the DS detector which was typically used for pointing at cluster
centroids. Each octant has at least a few × 100 counts to ensure validity of the χ2 statistic.
-- 10 --
Figure Captions
Figure 1a. Radial profile of the EUV surface brightness of A2199. A background as
determined from the data taken beyond the radius of 15 arcmin was subtracted.
Figure 1b. Radial profile of the DS Lex/B to PSPC R2 band ratio. The dotted line
represents the expected band ratio of ∼ 0.03 from a single temperature fit to the DS and
PSPC data, using the MEKAL code. The model for line-of-sight Galactic absorption is
Balucinska-Church & McCammon (1992); if the Morrison & McCammon (1983) model was
used instead, the predicted ratio would become 0.037. The EUV depletion within 2 arcmin
radius can be explained by the addition of an intrinsic absorbing column of ∼ 3.5 × 1019
cm−2. Elsewhere the rise of the ratio (and hence the degree of soft excess) with radius is
evident.
Figure 2a. DS Lex/B contours of the surface brightness of Abell 2199 (for details see text).
The peak emission is 1.98 × 10−3ph/arcmin2/s, and contours are labeled as percent of this
peak value, with the lowest contour at 8.4 × 10−5ph/arcmin2/s.
Figure 2b. PSPC R2 band contours of the surface brightness of Abell 2199 (for
details see text). The peak emission is 8.3 × 10−2ph/arcmin2/s. Next contours are
1.66×10−3, 2.8×10−3, 4.2×10−3, 7.2×10−3, 1.4×10−2, 2.9×10−2, and 7.4×10−2ph/arcmin2/s.
Figure 2c. Contours of DS Lex/B to PSPC R2 band ratio of Abell 2199. The dotted lines
mark regions where the DS brightness is 6 % above the background, and the contours are
labeled in units of 1 %. The expected value of this ratio was given earlier (see caption of
-- 11 --
Figure 1b).
Figure 3a. DS Lex/B contours of the surface brightness of Abell 1795 (for details see text).
The peak emission is 3 × 10−3ph/arcmin2/s, and contours are labeled as percent of this
peak value, with the lowest contour at 1.6 × 10−4ph/arcmin2/s (5 % above background).
Figure 3b. PSPC R2 band contours of the surface brightness of Abell 1795 (for
details see text). The peak emission is 9.5 × 10−2ph/arcmin2/s. Next contours are
8.6 × 10−4, 1.3 × 10−3, 2.1 × 10−3, 3.6 × 10−3, 6.7 × 10−3, 9.8 × 10−3, 1.4 × 10−2, 3 × 10−2 and
4.5 × 10−2ph/arcmin2/s.
Figure 3c. Contours of DS Lex/B to PSPC R2 band ratio of Abell 1795. The dotted lines
mark regions where the DS brightness is 25 % above the background, and the contours are
labeled in units of 1 %. The expected value of this ratio, based on a single temperature
ICM, is 1.8 % (Mittaz, Lieu, & Lockman 1998).
References
Arabadjis, J.S., Bregman, J.N., 1998, astro-ph/9810377.
Balucinska-Church, M. and McCammon, D., 1992, Astrophys. J. 400,
699 -- 700
Bowyer, S., Lampton, M., Lieu, R. 1996, Science, 274, 1338 -- 1340.
Bowyer, S., Lieu, R., Mittaz, J.P.D. 1998, The Hot Universe: Proc. 188th
IAU Symp., Dordrecht-Kluwer, 52.
Callanan, P.,J., Fabian, A.,C., Tennant, A., F., Redfern, R., M. and Shafer,
-- 12 --
R., A. 1987, M.N.R.A.S., 224, 781 -- 789.
David, L. P., Slyz, A., Jones, C., Forman, W. and Vrtilek, S.D. 1993,
Astrophys. J., 412, 479 -- 488.
Ensslin, T.A., Biermann, P.L. 1998, Astron. Astrophys., 330, 90 -- 98.
Ensslin, T.A., Lieu, R. and Biermann, P.L., submitted to Astron.
Astrophys., astro-ph 9808139.
Fabian, A.C. 1996, Science, 271, 1244 -- 1245.
Fabian, A.C. 1997, Science, 275, 48 -- 49.
Fusco-Femiano, R., Dal Fiume, D., Feretti, L., Giovannini, G., Matt, G.,
Molendi, S. 1998, Proc. of the 32nd COSPAR Scientific Assembly,
Nagoya, Japan (astro-ph 9808012).
Hwang, C. -Y. 1997, Science, 278, 1917 -- 1919.
Kaastra, J.S. 1998, Proc. of the 32nd COSPAR Scientific Assembly,
Nagoya, Japan .
Kaastra, J.S. 1992 in An X-Ray Spectral Code for OpticallyThin Plasmas
(Internal SRON-Leiden Report, updated version 2.0)
Lieu, R., Mittaz, J.P.D., Bowyer, S., Lockman, F.J., Hwang, C. -Y., Schmitt,
J.H.M.M. 1996a, Astrophys. J., 458, L5 -- 7.
Lieu, R., Mittaz, J.P.D., Bowyer, S., Breen, J.O., Lockman, F.J.,
Murphy, E.M. & Hwang, C. -Y. 1996b, Science, 274, 1335 -- 1338.
Lieu, R., Ip W.-I., Axford, W.I. and Bonamente, M. 1999, ApJL ,
510, 25 -- 28.
Mewe, R., Gronenschild, E.H.B.M., and van den Oord, G.H.J., 1985
Astr. Astrophys. Supp., 62, 197 -- 254
Mewe, R., Lemen, J.R., and van den Oord, G.H.J. 1986, Astr. Astrophys. Supp.,
65 511 -- 536
-- 13 --
Mittaz, J.P.D., Lieu, R., Lockman, F.J. 1998, Astrophys. J., 498, L17 -- 20.
Morrison, R. and McCammon D., 1983, Astroph. J., 270, 119 -- 122.
Mushotzky, R., Loewenstein, M., Arnaud, K.A., Tamura, T., Fukazawa, Y., Matsushita, K.,
Kikuchi, K., & Hatsukade, I., 1996, Astrophys. J., 466, 686.
Rephaeli, Y., Gruber, D. and Blanco, P. 1999, ApJL, 511, 21
Sarazin, C.L., Lieu, R. 1998, Astrophys. J., 494, L177 -- 180.
Sirk, M.M. et al 1997, Astrophys. J. Supp., 110, 347 .
Volk, H.J., Aharonian, F.A., Breitschwerdt, D. 1996, Space Sci. Rev., 75,
|
astro-ph/0303593 | 1 | 0303 | 2003-03-26T21:00:00 | Numerical Modeling of Gamma Radiation from Galaxy Clusters | [
"astro-ph"
] | We investigate the spatial and spectral properties of non-thermal emission from clusters of galaxies at gamma-ray energies between 10 keV and 10 TeV due to inverse-Compton (IC) emission, pion-decay and non-thermal bremsstrahlung (NTB) from cosmic-ray(CR) ions and electrons accelerated at cosmic shock and secondary e+- from inelastic p-p collisions. We identify two main emission region, namely the core (also bright in thermal X-ray) and the outskirts region where accretion shocks occur. IC emission from shock accelerated CR electrons dominate the emission in the outer regions of galaxy clusters, provided that at least a fraction of a percent of the shock ram pressure is converted into CR electrons. A clear detection of this component and of its spatial distribution will allow us direct probing of cosmic accretion shocks. In the cluster core, gamma-ray emission above 100 MeV is dominated by pion-decay mechanism and, at lower energies, by IC emission from secondary e+-. However, IC emission from shock accelerated electrons projected onto the cluster core will not be negligible. We emphasize the importance of separating these emission components for a correct interpretation of the experimental data and outline a strategy for that purpose. Failure in addressing this issue will produce unsound estimates of the intra-cluster magnetic field strength and CR ion content. According to our estimate future space borne and ground based gamma-ray facilities should be able to measure the whole nonthermal spectrum both in the cluster core and at its outskirts. The importance of such measurements in advancing our understanding of non-thermal processes in the intra-cluster medium is discussed. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 (2001)
Printed 29 October 2018
(MN LATEX style file v2.2)
Numerical Modeling of Gamma Radiation from Galaxy
Clusters
Francesco Miniati⋆
Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Str. 1, 85740, Garching, Germany
29 October 2018
ABSTRACT
We investigate the spatial and spectral properties of non-thermal emission from clus-
ters of galaxies at γ-ray energies. We estimate the radiation flux between 10 keV and 10
0-decay and non-thermal bremsstrahlung
TeV due to inverse-Compton (IC) emission, π
(NTB) from cosmic-ray (CR) ions and electrons accelerated at cosmic shocks as well as
secondary e± generated in inelastic p-p collisions. We identify two main region of pro-
duction of non-thermal radiation, namely the core (also bright in thermal X-ray) and
the outskirts region where accretion shocks occur. We find that IC emission from shock
accelerated CR electrons dominate the emission at the outer regions of galaxy clusters,
provided that at least a fraction of a percent of the shock ram pressure is converted into
CR electrons. A clear detection of this component and of its spatial distribution will
allow us direct probing of cosmic accretion shocks. In the cluster core, γ-ray emission
0-decay mechanism. At lower energies, IC emission
above 100 MeV is dominated by π
from secondary e± takes over. However, IC emission from shock accelerated electrons
projected onto the cluster core will not be negligible. We emphasize the importance
of separating the aforementioned emission components for a correct interpretation of
the experimental data and outline a strategy for that purpose. Failure in addressing
this issue will produce unsound estimates of the intra-cluster magnetic field strength
and CR ion content. According to our estimate future space borne and ground based
γ-ray facilities should be able to measure the whole non-thermal spectrum both in the
cluster core and at its outskirts. The importance of such measurements in advancing
our understanding of non-thermal processes in the intra-cluster medium is discussed.
Key words: acceleration of particles -- galaxies: clusters: general -- gamma rays:
theory -- methods: numerical -- radiation mechanism: non-thermal -- shock waves
1
INTRODUCTION
The structure and evolution of the large scale universe is a fundamental tool of cosmological investigation. Clusters of galaxies
in particular, being the largest bound objects in the universe, have been studied extensively in order to "weight" the cosmos and
to probe its structure. In the currently favored cold dark matter (CDM) hierarchical scenarios structure formation originates
from the collapse of primordial density perturbations seeded at an inflationary epoch (Peebles 1993). In this depiction, rich
clusters of galaxies correspond to linear perturbations on scales of order ∼ 10 Mpc comoving and, therefore, their potential
wells are expected to accommodate a representative fraction of the actual mass content of the universe (White et al. 1993).
By the same token, the environment there should be the end result of physical processes that occurred below or at that length
scale. And, in fact, observational evidence tells us that with respect to its pristine conditions the intra-cluster medium (ICM)
gas has been metal enriched by star formation processes, magnetized in some fashion and shock heated.
Perhaps spurred by observational progress, in recent years the subject of non-thermal process in, and their impact on,
the ICM has attracted much attention. In this regard, Volk et al. (1996) pointed out that metal enrichment by galactic
winds is likely accompanied by inter-galactic termination shocks where cosmic-ray (CR) acceleration might take place. In
addition, Ensslin et al. (1997) studied the case of CR protons escaping from of radio jets into the ICM and Kronberg et al.
(2001) that of black hole magnetic energy injection into the inter-galactic medium (IGM). The issue of CR ions produced
at cosmic shocks (Miniati et al. 2000) and their possible dynamical role was addressed by Miniati et al. (2001b) by means
⋆ [email protected]
2
F. Miniati
of a cosmological simulation of structure formation which included direct treatment of acceleration, transport and energy
losses of a CR component. They found that if shock acceleration takes place with some efficiency, CR ions, due to their long
lifetime against energy losses and to their efficient confinement by magnetic irregularities (Volk et al. 1996; Berezinsky et al.
1997), may accumulate inside forming structures and store a significant fraction of the total pressure there (Miniati 2000;
Miniati et al. 2001b).
In order to assess the level of non-thermal activity in groups and clusters direct observations of the associated non-
thermal emission is required. In particular, the integrated γ-ray photon flux above 100 MeV produced in the decay of neutral
π-mesons generated in p-p inelastic collisions, allows direct determination of the CR ion content and the pressure support
that they provide. For the sources of CRs mentioned in the previous paragraph, the authors that investigated them estimated
a γ-ray flux above 100 MeV that should be measurable by future targeted observations (Volk et al. 1996; Berezinsky et al.
1997; Ensslin et al. 1997; Blasi 1999; Atoyan & Volk 2000; Miniati et al. 2001b). However, in a recent paper Miniati (2002b)
computed the inverse Compton (IC) γ-ray emission from shock accelerated CR electrons scattering off Cosmic Microwave
Background (CMB) photons and for typical clusters of galaxies found it comparable to the γ-ray emission associated to
π0-decay. The importance of the former process was recently pointed out by Loeb & Waxman (2000) in the context of the
unexplained cosmic γ-ray background (see also Miniati 2002b; Keshet et al. 2003). According to results of Miniati (2002b), the
γ-ray flux from the leptonic component scales with the group/cluster temperature less strongly than its hadronic counterpart
(Miniati et al. 2001b). This means that it should dominate the γ-ray emission of small structures. In addition, as we shall see
in the ensuing sections, for conventional acceleration parameter values γ-ray from IC emission and π0-decay can be of the
same order even for a Coma-like cluster. Thus, the aimed detection of γ-ray emission from clusters may not necessarily reflect
the CR hadronic component and this should be borne in mind for a correct interpretation of the observational results.
With the objective of a correct diagnostic of future γ-ray observational results, in this paper we compute the radiation
spectrum and spatial distribution of non-thermal emission from groups/clusters of galaxies. We inspect the case of CR ions
and electrons accelerated at structure shocks as well as secondary electron-positrons (e±). However, we neglect the CRs that
might originate at galactic wind termination shocks (Volk et al. 1996) and radio galaxies. The computed spectrum extends
from hard X-ray (HXR) at 10 keV up to extreme γ-ray at 10 TeV. We examine the following emission processes: IC and
non-thermal bremsstrahlung (NTB) from shock accelerated electrons, IC emission from e± and γ-rays from π0-decay produced
by CR ions. We find that NTB emission is negligible throughout the spectrum. IC emission from shock accelerated electrons
should dominate at HXR energies and be comparable to the flux from π0-decay at γ-ray energies. Finally, IC emission from e±
is typically below these two components at all photon energies. However, we show that all of these three emission components
(i.e., e−, e± and π0) can potentially be individually measured by future γ-ray detectors with imaging capability by taking
into account their peculiar spatial distribution and spectral properties. In particular we find that IC emission from shock
accelerated CR electrons dominates the emission at the outer regions of galaxy clusters, provided that at least a fraction of
a percent of the shock ram pressure is converted into CR electrons. Therefore, a clear detection of this component and of its
spatial distribution will allow us direct probing of cosmic accretion shocks.
The paper is structured according to the following rationale: The numerical models for both the large scale structure and
the CR evolution are described in §2. The results on the non-thermal emission are presented in §3, discussed in §4 and finally
summarized in §5.
2 THE NUMERICAL MODEL
The objective of this study is to investigate the spatial and spectral properties of nonthermal γ-ray emission from CR electrons
and ions accelerated at large scale structure shocks surrounding clusters/groups of galaxies. The task at hand requires accurate
modeling of: (1) the large scale structure shocks where CR acceleration occurs; (2) the distribution within galaxy clusters
of the baryonic gas which provides the CR targets for production of π0 and secondary e±; and (3) the spatial propagation
and energy losses of the CRs. Furthermore, for a consistent description of the CR ions it is important to account for (the
acceleration due to) all the shocks that have processed the gas that ends up within the GCs themselves. That is because CR
ions with momenta above 1 GeV/c and up to ∼ 1018 eV/c have a lifetime against energy losses longer than a Hubble time.
This suggests that the formation of clusters and groups of galaxies be followed ab initio. Furthermore in order to reproduce
meaningfully the shocks where the CR acceleration occurs, this should be done within the "natural" framework of cosmological
structure formation. For this reason we have carried out a numerical simulations of structure formation which follows the
evolution of both the gas and dark matter in the universe starting from fully cosmological initial conditions. Details on the
cosmological simulation are given below in §2.1. In addition, with the numerical techniques described in the following sections
(§2.2) we were also able to include the evolution of CR ions, electrons and secondary e± by accounting explicitly for the
effects of CR injection (at shocks and through p-p inelastic collisions), diffusive shock acceleration, energy losses and spatial
propagation. Thus in addition to the gas distribution the simulation will also provide us with the information about both the
spatial and momentum distribution for each CR components. In particular this information will be used in the §3 to compute
the spatial and spectral distribution of nonthermal γ-ray emission within gravitationally collapsed structures. Some of these
results were already preliminarily presented in Miniati (2002a,c). Also, the simulation upon which the study is based, has
been used by the author to study the contribution of γ-ray emission from CRs in the diffuse intergalactic medium to the
cosmic γ-ray background (Miniati 2002b).
In the following subsections we describe in further detail the salient features of the employed numerical techniques. These
3
are basically the same as those used in previous related studies (Miniati et al. 2001b,a; Miniati 2002b,c), and the reader
familiar with them can skip directly to the result section in §3.
2.1 Large Scale Structure
The formation and evolution of the large scale structure is computed by means of an Eulerian, grid based Total-Variation-
Diminishing hydro+N-body code (Ryu et al. 1993). We adopt a canonical, flat ΛCDM cosmological model with a total mass
density Ωm = 0.3 and a vacuum energy density ΩΛ = 1 − Ωm = 0.7. We assume a normalized Hubble constant h ≡ H0/100
km s−1 Mpc−1 = 0.67 (Freedman 2000) and a baryonic mass density, Ωb = 0.04. The simulation is started at redshift z ≃ 60.
The initial density field is homogeneous except for perturbations generated as a Gaussian random field and characterized by a
power spectrum with a spectral index ns = 1 and "cluster-normalization" σ8 = 0.9. The initial velocity field is then computed
accordingly through the Zel'dovich approximation.
The choice of the computational box size is a compromise between the need of a cosmological volume large enough to
contain a satisfactory sample of collapsed objects and numerical resolution requirements. Ideally one would like to be able
to produce rich clusters of galaxies of the size of Coma cluster because they can be compared more easily with observations.
However, in order for such massive objects to develop out of the initial conditions described above, the size of the computational
box should be at least a few hundred h−1Mpc in size. Although this could be achieved with Adaptive Mesh Refinement
techniques, with the available computational resources and the code employed here the above box size would impose an
unacceptably coarse spatial resolution. Therefore the size of the computational box is set to L = 50 h−1Mpc. Given the
relatively small box size the dimensions of the largest collapsed objects, with a core temperature Tx ≃ 2 − 4 keV, are modest.
Therefore, when in §3.4 and §3.5 we make predictions for a Coma-size cluster (with Tx ≃ 8 keV) the computed simulation
results will be rescaled appropriately. The details of the rescaling procedure will be described in those sections.
Finally, the dark matter component is described by 2563 particles whereas the gas component is evolved on a comoving
grid of 5123 zones. Thus each numerical cell measures about 100 h−1kpc (comoving) and each dark matter particle corresponds
to 2 × 109 h−1M⊙.
2.2 Cosmic-rays
In addition to the baryonic gas and dark matter, the simulation also follows the evolution of three CR components, namely
CR ions and electrons injected at shocks and secondary e± generated in p-p inelastic collisions of the CR ions with the gas
nuclei. This is achieved through the code COSMOCR fully described in Miniati (2001, 2002b).
The dynamics of each of these CR component consist of the following processes: injection, diffusive shock acceleration,
energy losses/gains and spatial propagation. In this section we describe how each of these processes is modeled numerically
and how it is implemented in the simulation. We notice from the outset that CRs are treated as passive quantities, meaning
that their dynamical role is completely neglected both on the shock structure (test particle limit) and the gas dynamics.
2.2.1 Injection and acceleration at shocks
The first step in order to compute CR injection and acceleration at shocks, is to detect the shocks themselves. Here, as in
previous studies, shocks are identified as converging flows (∇ · v < 0) experiencing a pressure jump ∆P/P above a thresh-
old corresponding to a Mach number M=1.5. For the identified shocks both mass flux and Mach number are computed.
This is simply done by evaluating the jump conditions experienced by the flow as reproduced in the numerical simulation
(e.g., Landau & Lifshitz 1987). Shocks are found both around filaments and groups/clusters. In the latter case they show a
rather complex structure (see §3.1 and Miniati et al. 2000, for a detailed description of cosmic shocks). An example of gas
density distribution, velocity field and shock structure within a collapsed object is provided in Fig. 1, which illustrates the
dominance of asymmetry in the accretion flows and the existence of strong accretion shocks at the outskirts region followed
by weaker ones in the inner regions. A detailed description of Fig. 1 is delayed until §3.1 and we shall now focus on the
description of the adopted CR injection/acceleration scheme.
As we shall see in the following, the mass flux through the shock and the shock Mach number determine two important
quantities: respectively the CR injection rate and the shape of the accelerated distribution function. Injection and diffusive
acceleration at shocks take place on spatial scales of the order of the particles diffusion length which is
λd(p) = 1.1 (cid:16) E
GeV(cid:17)(cid:18) B
0.1µG(cid:19)−1
(cid:16)
us
102Km s−1(cid:17)−1
10−2 pc.
(1)
where E is the particle energy, B is the magnetic field strength, us the shock speed and we have assumed for simplicity Bohm
diffusion. As discussed in Miniati (2001, see also Jones et al. 1999) to properly follow injection and acceleration at shocks one
should at least resolve not only scales ∼ λd(p) but also the shock structure including the subshock. Because such small scales
can not be reproduced in the current simulation, injection and diffusive acceleration are not directly simulated but simply
modeled. In particular, given the mass flux across a shock, the assumed injection model determines the fraction of particles
that are converted into CRs. Then, the acceleration model determines how the injected CRs should end up distributed in
momentum space.
4
F. Miniati
For the injection of CR ions we follow the thermal leakage prescription (Ellison & Eichler 1984; Quest 1988; Kang & Jones
1995). That is, the post-shock gas is assumed to thermalize to a quasi-Maxwellian distribution function characterized by a
temperature Tshock. The ions/protons in the high energy tail of such distribution are assumed to be able to leak back upstream
and undergo the acceleration mechanism. The momentum threshold for such injection is set to pinj = 2c1mppkBTshock/mp,
where mp is the proton mass and c1 is a control parameter (Kang & Jones 1995). We adopt c1 > 2.5 which implies that a
fraction ∼ 10−4 of the particles crossing the shock be injected in the acceleration mechanism. In terms of shock ram pressure,
the pressure borne by the CR ions is a small fraction for weak shocks (M 6 3) and reaches 30 % for moderately strong shocks
(M ∼ 4 − 6). Given our simplified injection prescription for very strong shocks (M > 10) this fraction can approach unity or
so. Therefore, for consistency in these cases the fraction of injected particles is renormalized so that the CR pressure is always
limited below 40% of the shock ram pressure.
As for electrons (primary e−), we simply assume that the ratio between injected CR electrons and ions at relativistic
energies be given by a parameter Re/p. The introduction of this parameter simplifies the treatment of this process, which is
quite complex and yet to be fully understood (see, e.g., Ellison et al. 2000, for a discussion on this issue). Observationally, for
Galactic CRs this ratio appears to be in the range 1 × 10−2 − 5 × 10−2 (Muller & Tang 1987; Muller & et al. 1995; Allen et al.
2001). Also, based on EGRET observational upper limits on the γ-ray flux from nearby clusters (Sreekumar et al. 1996; Reimer
1999; Reimer et al. 2003), for a CR ion injection efficiency as assumed here Miniati (2002b) constrained Re/p 6 2 × 10−2.
Therefore, in the following we set Re/p = 10−2 and all of our results concerning primary e− are based on this assumption.
CR ions and electrons injected as described above, are quickly accelerated to energies much higher then the thermal
average through the diffusive shock acceleration mechanism. As already pointed out, given the prohibitively short length
scales over which the acceleration process takes place, its direct inclusion in the simulation is not possible. In addition, for a
typical shock with compression ratio close to four the acceleration time scale
τacc(p) = 21.1 (cid:16) E
GeV(cid:17)(cid:18) B
0.1µG(cid:19)−1
(cid:16)
us
102Km s−1(cid:17)−1
yr
(2)
is much shorter than the simulation dynamical time (computational time-step). Thus, as in previous studies, we assume that
the injected CRs are accelerated instantly and distribute them in momentum according to a power-law distribution extending
from injection up maximum energy, i.e.,
f (p) = f0(cid:18) p
pinj(cid:19)−q
,
pinj 6 p 6 pmax.
(3)
For CR ions f0 is determined through the aforementioned thermal leackage injection model and for CR electrons it is such that
at relativistic (> GeV) energies their ratio to accelerated protons is ≃ Re/p. According to the test particle limit adopted here,
the log-slope of the distribution function is related to the shock Mach number, M , as q = 3(γ +1)/[2(1−M −2)] = 4/(1−M −2)
for γ = 5/3. This implies relatively flatter distribution functions for stronger shocks.
Based on cluster/group properties, we found it appropriate to follow CR ions between momenta pmin = 10−1GeV/c
and pmax = 106GeV/c. In fact, CR ions below pmin quickly loose energy due to Coulomb losses and beyond 107 GeV/c
their confinement within cosmic structures becomes difficult (Volk et al. 1996; Colafrancesco & Blasi 1998). CR electrons are
followed between pmin = 15 MeV and 2 × 10 TeV. The latter is a reasonable value for the maximum energy to which CR
electrons are shock accelerated, provided a magnetic field of order 0.1µG.
2.2.2 Secondary Electrons and Positrons
In addition to ions and electrons injected and accelerated at shocks, CR electrons and positrons are also produced in hadronic
collisions of CR ions with the nuclei of the intra-cluster gas. The parent CR ions are those computed in the simulation.
Secondary e± are generated in the decay of charged muons according to
µ± → e± + νe(¯νe) + ¯νµ(νµ)
(4)
which in turn are produced in the following reactions
p + p → π± + X,
p + p → K ± + X,
π± → µ± + ν(¯ν)
K ± → µ± + ν(¯ν)
K ± → π0 + π±
(5)
(6)
(7)
Additional e± are produced in cascades analogous to these by the following interactions: p+He, α+H and α+He. These are
included by assuming a helium number fraction of 7.3% for the background gas and a ratio (H/He) ≃ 15 at fixed energy-
per-nucleon for the CRs (Meyer et al. 1997). The distributions of the e± thus generated are obviously related to those of the
parent CR ions. In particular, if the latter are distributed according to a power law, so are also the former.
2.2.3 Energy losses and spatial propagation
The numerical treatment of the CR dynamics is completed by accounting for both spatial transport and energy losses/gains
as described by the diffusion-convection equation (e.g., Skilling 1975a,b). In order to compute a numerical solution to this
5
equation, we define a grid in momentum space by dividing it in Np logarithmically equidistant intervals (momentum bins). For
each mesh point of the spatial grid, xj, the CR distribution function is then described by the following piecewise power-law
(Jones et al. 1999; Miniati 2001)
f (xj, p) = fj (xj) p−qi(xj ),
1 < pi−1 6 p 6 pi,
(8)
where pi... are the bins' extrema. Spatial propagation and energy losses of the accelerated CRs are then followed by solving
numerically a diffusion-convection equation that has been multiplied by '4π p2' and integrated over each momentum bin.
Written in the comoving coordinates system adopted for the other simulated hydrodynamic quantities, such equation reads
(Miniati 2001)
∂n(xj, pi)
∂t
= −
1
a
∇· [u n(xj, pi)] +nh(cid:16) a
a
+
1
3a
∇ · u(cid:17) p + bℓ(p)i 4πp2f (xj, pi)opi
pi−1
+ Q(xj, pi).
(9)
pi−1
Here n(x, pi) = R pi
4π p2 f (xj, p) dp is the comoving number density of CR in the i-th momentum bin and Q(xj, pi) is
a comoving source term, i(xj, p), describing either shock injection or e± generated in p-p collision and integrated also over
the i-th momentum bin. In addition, a is the cosmological expansion factor, a/a = H(z) is the Hubble parameter defining
the cosmic expansion rate. Thus the first part of the second term in eq. (9) describes adiabatic losses/gains. Finally, bℓ(p)
represents energy losses due to Coulomb collision, bremsstrahlung, IC and synchrotron emission for electrons and e±; and
Coulomb and p-p inelastic collision for ions. All of them are fully described in Miniati (2001). In eq. (9) u is the velocity field of
the baryonic gas to which the CRs are assumed to be coupled through a background magnetic field. Thus once accelerated the
CRs are passively advected with the gas flow. Notice that eq. (9) does not include the diffusion term present in the ordinary
diffusion-convection equation. This is because, as pointed out in Miniati (2001, see also Jones et al. 1999) for typical flows in
the simulation with u ∼ a few 100 km s−1 diffusion over spatial scales ∆x ≃ 100 kpc is much slower than advection and can
be neglected away from shocks. This is true for values of the diffusion coefficients typically assumed in the literature (see, e.g.,
Volk et al. 1996; Berezinsky et al. 1997). The value inferred for these diffusion coefficients is based on the assumption that
the magnetic field fluctuations are described by a turbulent spectrum and that the total magnetic field energy density is that
characteristic of µG strong magnetic fields.
Integrating eq. (9) is sufficient for an accurate treatment of the ionic component. In this case, after advancing the solution
to eq. (9) one time-step, both fj (xj) and qi(xj) defining the CR distribution function in eq. (8) are computed based on the
updated values of n(xj, pi) and by assuming that f (xj, p) is continuous at each bin interface.
ε(x, pi) = 4πR pi
For CR electrons, however, severe energy losses and distribution cutoffs render their numerical treatment much more
delicate. For this reason for this component in addition to n(x, pi) we also follow the correspondent bin kinetic energy
p2 f (x, pi) T (p) dp. Here T (p) = (γ − 1) me c2 is the particle kinetic energy and γ = [1 − (v/c)2]−1/2 is the
Lorentz factor. The equation describing the evolution of ε(xj, pi) is obtained in analogy to eq. (9), by integrating over the i-th
momentum bin the same diffusion-convection equation (more properly a kinetic equation) that has been multiplied by T (p).
In comoving units this reads (Miniati 2001)
pi−1
dp + S(xj, pi),
(10)
∂ε(xj, pi)
∂t
= −
1
a
∇· [u ε(xj, pi)] +(cid:2)b(p) 4π p2 f (xj, p) T (p)(cid:3)pi
pi−1
−Z pi
pi−1
b(p)
4πcp3f (xj, p)
pmec2 + p2
pi−1
where S(xj, pi) = 4π R pi
i(xj, pi) p2T (p) dp, b(p) now includes both adiabatic loss terms and those described by bℓ(p), and
the third term on the right hand side includes contributions from CR pressure work and sink terms. The evolution of the CR
electrons is obtained by integrating numerically eq. (9) and (10) through a semi-implicit scheme described in Miniati (2001,
2002b). The slope of the distribution function at each grid point and momentum bin is then determined self consistently by
the values of n(xj, pi) and ε(xj, pi).
However, for the simulated CR electrons and e± with momenta between 102 GeV/c and 20 TeV cooling time due to
IC losses, τcool ∼ 105 − 107yr, is much shorter than the typical computational time-step, ∆t ∼ 107 − 108yr. Thus, because
uτcool ≪ u∆t 6 ∆x (Courant condition), these particles only exist in grid cells where the source term i(xj, p) 6= 0 and will not
propagate outside it. Thus, as discussed in (Miniati 2002b), in this case it is computationally advantageous and numerically
correct to take directly the steady state solution to eq. (9) as
n(xj, pi) = 4πZ pi
pi−1
p2 fc(xj, pi) dp = −4πZ pi
pi−1
dp
b(p) Z ∞
p
ρ2 i(xj, ρ) dρ
(11)
[a similar expression holds for eq. (10)]. Notice that n(xj, pi) refers to the cell-volume averaged number density of CRs.
Physically, the expression in eq. (11) represents a summation of all the individual populations of CR electrons within the grid
zone xj that, as they emerge from the acceleration region, and are being advected away from the shock, develop a cut-off due
to energy losses. For what follows in the next sections it is relevant to notice that for an injection spectrum i(p) ∝ p−s and
for the case of energy losses dominated by IC emission, b(p) ∝ p2, the steady state distribution given in eq. (11) implies a
cell-volume averaged distribution function hf (p)i ∝ n(xj, pi)/∆x3 ∝ p−(s+1).
Finally the CR ion distribution function is mapped by 4 momentum bins. For the electrons and e± we used 5+1 momentum
bins. The first five are logarithmically equidistant and cover a momentum range between pmin = 15 MeV up to p2 = 102
GeV. The last electron momentum bin stores the CR electron steady-state distribution function between p2 = 102 GeV and
6
F. Miniati
Figure 1. Color image showing the density distribution on a plane passing through the center of the collapsed object. It is in units of
cm−3 and the various level correspond to values indicated by the colorbar on the right of each panel. Arrows describe the velocity field
on the same plane. Their number has been reduced by a factor nine for clarity purposes. Finally, blue contours indicate the isolevel
of compression (∇ · v <0) where shoks occur. Narrow countour features correspond to location of strong shocks. Left and right panel
correspond to X-Y and Z-X coordinate planes respectively. Axis scales are measured in h−1Mpc.
20 TeV given by eq. (11). Our tests indicate that increasing the number of momentum bins for either CR component (ions
and electrons) does not affect the results.
3 RESULTS
In the following sections we present results based on the simulation just described on the emission of high energy radiation
from CRs in groups/clusters of galaxies. In particular, we inspect the spectra, radial dependence and synthetic images for
various emission mechanisms and for each computed CR population. For the purpose we excise out of the simulation box
regions containing individual virialized objects. For each of them we compute the nonthermal radiation produced through
emission processes described below, by using the simulated hydro (density) and CR (distribution functions) quantities. For
simplicity in the following we focus only on a couple of simulated virialized objects, although the studied radiation properties
are independent of our peculiar choice. Finally, in §3.2 and §3.3 we describe the qualitative properties of the nonthermal
radiation whereas in §3.4 and §3.5 we will attempt quantitative predictions for a Coma-like cluster of galaxies.
3.1 Density and Velocity Structure
First we consider the largest virialized object in the computational box. This is characterized by an X-ray core temperature
Tx ≃ 4 keV. The object is partially described in fig. 1 where the two panels show two dimensional cuts across the coordinate
planes X-Y and Z-X respectively of the density distribution (in color scale), the velocity field (arrows) and the location of
shocks (contours). The gas density is nc ∼ 6 × 10−3 cm−3 at the core and drops by about 2 orders of magnitudes at a
distance of a few Mpc. The tree-dimensional velocity characterizing the accretion flows is typically of order of 103 km s−1.
The velocity field converges toward the mass concentration and becomes quite complex close to it. As described in more detail
in Miniati et al. (2000), this is due to the large asymmetry of accretion flows. Noticeably, stream along filaments, which carry
more momentum, reach closer to the central regions before being shocked. As a result the ensuing shock structure is also
complex. The largest shocks are found up to ∼ 5 h−1Mpc from the core of the collapsed object. Additional shocks are present
at intermediate distances, especially along filaments as already pointed out. These shocks have Mach numbers that range from
up to ∼ 100 in the most outlining regions, to 3 − 10 along filaments. Close to the high temperature core some discontinuities
are also found. However, as it is apparent from the broad features of the isocontours, these are weak, tran-sonic shocks.
For the selected object the energy in CR ions is ECR ≃ 6 × 1061 erg in the inner 1.5 Mpc and corresponds to about 26%
of the thermal pressure within the same volume. We notice the presence in the neighborhood of the collapsed objects of a
small structure which is is marginally visible above and to the right of it, in the left and right panels of fig. 1 respectively. As
we shall see in the following, this will affect (particularly, will increase) the computed radiation spectrum at large radii.
3.2 Radiation Spectrum
In fig. 2 the quantity 'ε Lε' in units 'keV s−1' associated with the selected object is plotted as a function of photon energy, ε.
'Lε' is the volume integrated spectral power and is defined as
Lε(ε, R) = ZV(R)
dV jε(ε)
(12)
where jε(ε) is the spectral emissivity in units 'keV cm−3 s−1 keV−1' and V (R) = 4πR3/3 is an integration volume defined
by a radius R. Here we set R = 5 h−1Mpc in order to account for the shock accelerated electrons which are typically located
away from the cluster center (see fig 1). We consider the following emission processes: π0-decay (solid line), IC emission
from shock accelerated electrons (dotted line) and from secondary e± (dashed line) scattering off CMB photons, and NTB
from shock accelerated electrons (dot-dashed line). Thermal X-ray photons, of energy εX , can be neglected as seeds for IC
emission of γ-rays with respect to CMB photons to first order approximation. In fact, although the number of available IC
scattering electrons is larger by a factor (hεXi/hεCM Bi)(q−3)/2 this is not sufficient to compensate for the lower number of
seed photons (especially for flat CR distributions such as those produced at accretion shocks) and for the IC cross-section
suppression due to Klein-Nishina effects which enter for x ≃ 1.7[ε(GeV)εX (keV)]1/2 > 1 (see also, e.g., Ensslin & Biermann
1998; Berrington & Dermer 2002).
According to fig. 2, both emission from π0-decay and IC from primary e− contribute significantly to the γ-ray emission
above 100 MeV. The actual proportion between these two components is determined by two factors: (a) The ratio of accelerated
electrons and ions at relativistic energies, which has never been measured for cluster shocks. It is represented by the parameter
Re/p and was set to 10−2 in the current study. (b) The temperature of the group/cluster, from which the γ-ray luminosity
7
Figure 2. Radiation spectrum extending from 10 keV up to 10 TeV produced by the following emission processes: IC (dotted) and NTB
(dash-dotted) from shock accelerated CR electrons, IC emission from e± (dashed) and γ-rays from π0-decay produced by CR ions (solid).
from π0-decay and IC emission depend differently. In fact, for both processes the temperature dependence is expected to be
of power-law type, but with a moderately steeper index for the hadronic emission component (Miniati et al. 2001b; Miniati
2002b). Thus the relative strength of these two contributions as presented in fig. 2 is only indicative at this stage. However, it
is within the objectives of the present paper to investigate possible strategies for their experimental determination. At HXR
and soft γ-ray energies below 10 MeV, IC from primary e− dominates the spectrum although the contribution from e± is not
negligible. However, as we shall see in the next sections, this source of emission should also be detectable at HXR energies
owing to its different spatial distribution with respect to the primary e− .
Noticeably the radiation spectra of these three components are rather "flat". This indicates that the primary CR ions
and e−'s were generated in shocks of at least moderate strength, i.e., M > 4. Indeed, for the hadronic component, the γ-ray
emissivity from a CR ion distribution f (p) ∝ p−qp , is jε(ε) ∝ ε−(4qp −13)/3 (cf. Mannheim & Schlickeiser 1994). Thus, the
case of an emission spectrum such that εLε ∝ εjε ∝ ε0, implies a CR ion distribution with qp ≃ 4. As for both primary
electrons and secondary e± , because of the severe energy losses suffered by these particles, the emission spectrum must be
related to their steady state distributions. As pointed out at the end of §2.2, this is characterized by a steady-state log-slope,
qss = qsource + 1, that is steeper by one unit with respect to the source spectrum. The produced IC emissivity is of the
form jε(ε) ∝ ε−(qss−3)/2 which satisfies the condition εjε ∝ ε0 for qss ∼ 5 and, therefore, qsource ∼ 4. For primary e− this
straightforwardly requires that these CRs were accelerated at strong shocks. For secondary e± with a source term of the form
fe± (εe± ) ∝ ε
(cf. Mannheim & Schlickeiser 1994), it requires that the parent CR ions were accelerated at strong
shocks and that their log-slope distribution be qp ≃ 4.
−4(qp −1)/3
e±
Finally, NTB is negligible for all purposes. Notice that latter component includes also the contribution from trans-
relativistic electrons. These manage to propagate away from the acceleration region (i.e., a shock) for a short time before
being re-absorbed into the thermal pool due to Coulomb losses. This trans-relativistic component enhances the NTB emissivity
with respect to that produced by the bulk of the relativistic electrons. In fact, it dominates the emission below 10 MeV causing
a sort of bump in the shape of the spectrum about this photon energy (dot-dashed line in fig. 2).
3.3 Luminosity Volume Dependence
We now inspect the spatial distribution of the emissivity associated with the various processes presented in the previous
section. As anticipated there, such information allows in principle a measurement of each individual emission component
(except NTB). With this in mind, in fig. 3 we show the radial dependence of the integrated photon luminosity above 100 keV,
that is
Lγ (> 100 keV, R) = Z>100 keV
Lε(ε, R) dε
(13)
where R, the radius defining the integration volume, is now a varying parameter. Obviously, at these low energies there is no
contribution from the CR ionic component. Thus the diagram contains only three curves illustrating IC emission from primary
8
F. Miniati
Figure 3. Radial dependence of the integrated photon luminosity above 100 keV for the following emission processes: IC (dotted) and
NTB (dash-dotted) from from shock accelerated CR electrons, IC emission from e± (dashed).
Figure 4. Radial dependence of the integrated photon luminosity above 100 MeV for the following emission processes: IC (dotted) and
NTB (dash-dotted) from shock accelerated CR electrons, IC emission from e± (dashed) and π0-decay (solid).
e− (dotted) and secondary e± (dashed) and, for completeness, NTB from e− (dot-dashed). This plot clearly shows how the
radiation emitted by primary e− and secondary e± originates in spatially separate regions. In fact, the latter component
saturates quickly within the central Mpc and dominates the total photon production there. On the contrary, the contribution
from the former component becomes significant only at a distance of about 1.5 h−1Mpc from the cluster center and keeps
increasing up to several Mpc from there.
The situation is analogous at γ-ray energies. This is illustrated in fig. 4 where the integrated photon luminosity above
100 MeV, Lγ (> 100 MeV, R) - defined by an equation analogous to eq. (13) - is plotted versus radial distance. Here it is
9
Figure 5. Synthetic map of the integrated photon flux above 100 keV (top) and 100 MeV (bottom) in units "ph cm−2 s−1 arcmin−2"
from IC emission by secondary e± (top-left), primary e− (top-right, bottom-right), and π0-decay (bottom-left). Each panel measures 15
h−1Mpc on a side.
the emission from π0-decay (solid line) that saturates at relatively short distances from the cluster center and dominates the
photon emission in the inner regions. As before, IC emission from primary e− (dotted line) reaches a substantial level only
outside a distance ∼ 1.5 h−1Mpc. IC emission from e± (dashed line) is now unimportant and, as in the previous case, NTB
from primary e− (dot -- dashed line) is completely negligible.
The radial dependence of the various radiation components presented in fig. 3 and 4 reflects both the spatial distribution
of the emitting particles as well as the nature of the emission process. Thus, on the one hand, both e± and π0 are produced at
the highest rate in the densest regions where both the parent CR ions and target ICM nuclei are most numerous. Consequently,
e± IC and π0-decay emissivities are strongest in the cluster inner regions and quickly fade toward its outskirts. Notice that in
the case of fig. 3 and 4 this behavior is slightly altered by the presence of a nearby object (similar to the case in fig. 5 below)
which causes these integrated photon luminosities to slowly grow even after few core radii instead of leveling out.
On the other hand, because of severe energy losses, the emitting high energy primary e− are only found in the vicinity of
strong shocks where they are being accelerated. Notice that in this case the shocks must be strong so that enough particles
are accelerated to high energies to produce appreciable emission. Thus, because the strongest shocks are located at the cluster
outskirts rather than at its core where the ICM temperature is already high (fig. 1; also cf. Miniati et al. 2000), the spatial
distribution of the emissivity is now reversed with respect to the previous case. Notice that, in line with this analysis, the
drop of IC emissivity (from primary e− ) toward the inner regions is slightly more abrupt at γ-ray energies (fig. 4) than at
HXR energies (fig. 3).
3.4 Synthetic Maps
In this section we present synthetic maps produced by non-thermal emission processes. These maps were obtained upon
simple integration along the line of sight of the photon emissivity in the thin plasma approximation. We use them to study
10
F. Miniati
Figure 6. Left, Right: Synthetic map of the total photon flux above 100 keV (left) and 100 MeV (right) in units 'ph cm−2 s−1 arcmin−2'.
Left panel includes IC emission from primary e− , secondary e± and NTB. Right panel includes emission due to π0-decay , IC from
primary e− , secondary e± and NTB (c). Center: Synthetic map of the bolometric X-ray emission from thermal bremsstrahlung in units
'erg cm−2 s−1 arcmin−2'. Each panel measures 15 h−1Mpc on a side.
the morphology of the emitting region. In addition we investigate which emission component contributes most to the radiation
spectrum when the latter is extracted from different spatial regions. Unlike plots in fig. 3 and 4, using synthetic maps will
automatically account for projection effects.
We will attempt to relate the present analysis to the case of Coma cluster which, for its large size and relative vicinity,
is likely the best source candidate for the detection of non-thermal, high energy radiation. Our calculations are now based
on a different collapsed object extracted as before from the simulation box but relatively more isolated (and therefore better
suited for the current purpose) than the previous one. It has a temperature of a few keV, a core density of 2 × 10−4 cm−3.
11
This object is smaller than Coma cluster and for that reason we will renormalize the emission components as follows.1 As
for the hydro quantities we retain the temperature and density profiles produced in the simulation but renormalize them by
(multiplying each of them by) the ratio of the observed and simulated core values. The following observed values for Coma
cluster are taken: a temperature Tx = 8.2 keV (Arnaud et al. 2001) and a gas density of ngas ≃ 3 × 10−3 cm−3. The thermal
energy within 1 Mpc computed according to temperature and density thus renormalized corresponds to 1.6 × 1063 erg.
For the IC emission from primary e− we have used the recent finding of Miniati (2002b) indicating that the γ-ray flux due
to this process scales with the X-ray temperature as Fγ ∝ T 2.6
x . The total number of secondary e±, Ne± , and the associated
IC flux F ±
IC , are normalized by assuming that these particles are responsible for producing Coma radio halo through emission
of synchrotron radiation. For a given measured radio flux, Ssy, and an assumed average ICM magnetic field strength, hBi,
the total number of emitting particles scales as
F ±
IC ∝ Ne± ∝
Ssy
hBi1+α ,
(14)
where α ≃ 1 is the radiation spectral index. The assumption that Coma radio halo is produced by secondary e± , as opposed
to high energy e− of different origin, allows us to also fix the total number of parent CR ions, Ncri, and the ensuing π0-decay
fluxes, Fπ0→γγ . In fact, since the steady-state total number of secondary CR electrons accords to
Ne± ∝ Ncrihρgasi
1
1 + UB/UCMB
based on this and eq. (14) one finds
Fπ0→γγ ∝ Ncrihρgasi ∝ Ssy
1 + UB
UCMB
hBi1+α ,
(15)
(16)
where UB = hBi2/8π and UCMB are the energy density in magnetic field and CMB radiation field, respectively. For the
synchrotron flux needed in (14) and (16) we adopted the value S1.4GHz = 640 mJy measured at 1.4 GHz by Deiss et al. (1997).
When extracting synthetic spectra in the next section we consider two cases for the magnetic field strength corresponding
to hBi ∼ 0.15µG and hBi ∼ 0.5µG (Kim et al. 1990). The former choice implies a ratio of CR ions to thermal energy about
30 % as opposed to almost 4% for the latter. This second case is assumed to produce the synthetic maps described below.
Notice that its "renormalized" values would imply a lower CR ion acceleration efficiency then described in §2.2.1 and would
also correspond to a parameter Re/p ∼ 0.05.
It is worth pointing out that according to eq. (14) and (16) smaller values of hBi than assumed here would imply corre-
spondingly higher fluxes of IC emission and γ-ray from π0-decay . For very weak fields, then, most of the assumed radio emitting
particles could not be of hadronic origin due to EGRET upper limits on γ-ray flux from Coma cluster (Blasi & Colafrancesco
1999; Miniati et al. 2001a, cf. also fig. 8). In the opposite limit of stronger magnetic fields than assumed here, less emitting
particles would be required to produce the same radio emission. As a consequence, the associated IC flux is reduced propor-
tionally to the inverse of the magnetic energy density [cf. eq. (14)]. However, as indicated by the expression in eq. (16), for
α ≃ 1 and hBi ≫ pUCMB/8π ∼ 3.3 µG, the γ-ray production from π0-decay reaches only a floor value. This is because even if
the higher magnetic field implies a smaller number of radio emitting e± particles, in steady state a minimum rate of hadronic
interactions is necessary to compensate for the increased synchrotron losses. In fact the enhanced synchrotron emission rate
from the total particle distribution (∝ hBi1+α ∼ hBi2) is counterbalanced by the reduced number of steady state emitting e±
[∝ Ne± ≃ 1/hBi2 - see eq. (15)]. Thus in secondary models of radio halos, there is a lower limit to the expected γ-ray flux
from π0-decay .
In fig. 5 we present synthetic maps of the integrated photon flux above 100 keV (top panels) and 100 MeV (bottom
panels) for an assumed hBi ≃ 0.5µG. Right panels are associated with IC emission from primary e− whereas left panels
correspond to emission due to IC from secondary e± (top-left) and π0-decay (bottom-left). The top and bottom panels of fig.
5 are combined in the left and right panels of fig. 6 respectively to produce a synthetic map of the total photon flux above
100 keV and 100 MeV. (For completeness these maps also include negligible contributions from the additional processes that
were investigated in the previous section, e.g., NTB.) A synthetic map associated with the (not rescaled) bolometric X-ray
emission from thermal bremsstrahlung is also shown in the central panel of fig. 6 to allow for comparison between thermal and
non-thermal processes. Non-thermal maps are presented in units of "ph cm−2 s−1 arcmin−2" whereas thermal X-ray surface
brightness is in units "erg cm−2 s−1 arcmin−2".
The synthetic images in fig. 5 show that the non-thermal emissivity is remarkably extended. In accord with the previous
section we find that the emission from π0-decay (bottom left) and e± (top left) is confined to the cluster core. There it creates
a diffuse halo which rapidly fades with distance from the center. Also, as to be expected, there is a strong correlation between
the spatial distribution of both these emission components and that of the thermal X-ray emission (fig. 6). On the other hand,
IC emission from primary e− is distributed over a much more extended area. Moreover, it is characterized by a strikingly
rich and irregular structure. This is a direct reflection of the complex "web" of shocks that reside at the outskirts of galaxy
clusters first pointed out by Miniati et al. (2000). The morphology of the emissivity in fig. 5 and 6 is quite similar, although
noticeably the central diffuse emission stands out more prominently in the high energy γ-ray map.
1 Notice that this "renormalization" procedure can be applied to any other cluster for which sufficient observational data are available.
12
F. Miniati
Figure 7. Synthetic spectra extracted from the core region (top) and the outskirts region (bottom).
3.5 Synthetic Spectra
In order to assess the separability of the different emission components, in fig. 7 we have produced synthetic spectra taken
from a core (top) and an outskirts region (bottom). The extension of the core region corresponds to an angular size of 1o (or
2 Mpc diameter at the red-shift of Coma cluster) whereas the outskirts region is defined as an annular ring with inner and
outer radii of 0.5o and 1.5o (or 1 Mpc and 3 Mpc at the red-shift of Coma cluster) respectively. For the core region we consider
two values for the magnetic field strength (0.15 and 0.5 µG). For each emission region in fig. 7 we plot the predicted flux
contribution due to the main emission processes considered so far namely, IC emission from primary e− (dot) and secondary
e± (dash and dot -- dash), and π0-decay (solid and long -- dash). Notice that as the magnetic field strength is increased from 0.15
to 0.5 µG the fluxes due to IC from e± and π0-decay drop by about a factor ∼ 10 (cf. eq. 14 and 16). Fig. 7 shows that at
high photon energy (> 100 MeV) the spectrum of radiation is indeed dominated by π0-decay in the core region (top panel)
and by IC emission from primary e− in the outer region (bottom). In the latter case, the residual π0-decay component is
actually due to the presence in the field of view of a small structure north-west of the selected object (see fig. 5; since this is
not part of the main cluster, we do not renormalize it according to various magnetic field values as it was done for the core
emission). In principle any contribution such as this can be removed by excision of emission regions associated with thermal
X-rays from structures appearing in the field of view. Below ∼ 100 MeV the flux from the outskirts region is still strongly
dominated by IC emission from primary e− . The contribution from the latter is much reduced in the narrower field of view
of the core region. However, for the larger field case it is still significant and at the level of IC emission from secondary e± .
The relative amount of radiation flux from these two components further depends on the actual shock structure subtended
by the field of view. Nevertheless the point is that for magnetic field of a few tenths of µG they are expected of comparable
intensity. In order to separate them out one could measure the radiation flux as a function of radial distance in the outer
regions, where it presumably is only due to CR electrons accelerated at accretion shocks, and then extrapolate its value for
the core region. For this purpose, an angular resolution ∼ 10′ or so should be sufficient.
Finally, above the spectra are characterized by a cutoff which appears at photon energies about 1 TeV. For IC emission
from primary e−, for which this feature is sharpest, this is mostly due to the maximum momentum, pmax = 20 TeV, of the
accelerated particles. For the other processes, it is rather caused by absorption due to the reaction γγ −→ e+e−. In fact, pair
production becomes important at TeV energies due to the presence of diffuse background radiation field at optical/infrared
wavelengths. The spectra in fig. 7 were thus corrected through a factor exp(−τγγ ) and by assuming an optical depth for pair
production (e.g., Coppi & Aharonian 1999)
τγγ (εγ ) ≃ 0.14 (cid:16) εγ
1 TeV(cid:17) (cid:18)
u(ε∗)
2 × 10−3eV cm−3(cid:19) (cid:16) zComa
0.023 (cid:17) h−1
(17)
where ε∗ = 4m2
from Dwek & Arendt (1998).
ec4/εγ is the target photon energy and u(ε∗) is the energy density carried by the background radiation taken
13
Figure 8. Total radiation spectra extracted from the same spatial regions illustrated in fig. 7 -- core for two different magnetic field
values (dot and long -- dash), outskirts (short dash) and their summation (solid and dot -- dash) -- compared to nominal sensitivity limits of
future γ-ray observatories (thick-solid lines). For INTEGRAL-IBIS imagers (ISGRI and PICsIT) the curves correspond to a detection
significance of 3σ with an observing time of 106 s. All other sensitivity plots refer to a 5σ significance. EGRET and GLAST sensitivities
are shown for one year of all sky survey whereas for Cherenkov telescopes (MAGIC, VERITAS and HESS) for 50 hour exposure on a
single source.
4 DISCUSSION
In fig. 8 we plot sensitivity limits of planned γ-ray observatories together with the total radiation spectra from the core (dotted
line) and outskirts (dashed line) regions that were presented in fig. 7. The sensitivity limits are plotted for point sources. For
extended sources it will be worse by roughly a factor that goes as θsource/θinst that is the ratio between the source size and the
angular resolution of the instrument. For Cherenkov telescopes θinst ∼ 0.1o In any case, according to this plot, several future
experiments should be sensitive enough to detect the computed non-thermal emission at most photon energies. In particular
the IBIS imager onboard INTEGRAL should readily measure the flux between 100 keV and several MeV. In addition GLAST
and Cherenkov telescopes should be able to detect both the core γ-ray emission from π0-decay as well as the IC flux directly
produced at accretion shocks above 100 MeV and 10 GeV respectively. One should notice that resolving sharp and complex
structures such as those reported in fig. 5 and 6 is a real challenge, especially for coded mask techniques employed by several
current and planned γ-ray imagers. Although a nominal resolution of ∼ 10′ in principle should be sufficient to identify some
of the bright shock features, the final outcome of such measurements will hinge on the actual source fluxes and instruments
performance.
The measurement of the non-thermal radiation spectra at several photon energies spanning the range illustrated in fig.
7 provides important information about the physical conditions in clusters. First, besides the mere detection of HXR and
γ-ray radiation from clusters, important information is embedded in the spatial distribution of their surface brightness. In
particular the extended emission component corresponds to the location of accretion shocks. Merger shocks have occasionally
been observed in the core of clusters as relatively weak temperature jumps. However, strong accretion shocks have yet to
be observed directly. Thus, detection and imaging of IC emission from primary electrons would provide an opportunity to
directly observe these accretion shocks given their wide angular extension and their unique morphology.
In addition, the flux about 100 keV will give us the first direct estimate of the energy density of CR electrons and, together
with radio measurements, will allow an estimate of (or upper limits on) the ICM magnetic field. In this respect for a correct
interpretation of the data it will be necessary to account properly for the fluxes contributed by both the CR electrons in the
core and those directly accelerated at the outlining shocks. In fact, the former likely propagate in relatively highly magnetized
regions and produce a substantial radio emission. On the other hand, because of the expected decline of the magnetic field
strength toward the outer regions, the latter might generate only a weak radio emission but, nonetheless, a strong IC flux.
Since, as shown in fig. 7, the HXR flux produced by this second component can easily dominate the total flux at this spectral
range (the details will depend on the assumed normalizations on acceleration efficiency and average magnetic field strength)
separating the contributions from CRs in the cluster core and external accretion shocks will be an important step for correctly
measuring ICM magnetic fields. A similar point was qualitatively discussed already in Miniati et al. (2001a) and was also
14
F. Miniati
addressed in Brunetti et al. (2001) in the context of a their "multiphase" acceleration model. In particular Brunetti et al.
(2001) pointed out that the magnetic field in the outer region, where most of the IC emission is produced due to the larger
emitting volume, can be much lower than in the core where the radio emission originates. In that model all the radio and
HXR emitting particles are accelerated with the same mechanism, whereas in our model the CR electrons that generate the
HXR flux are accelerated in the outer accretion shocks and never make it to the cluster core.
It is worth mentioning that our estimated non-thermal flux from IC emission above 20 keV is quite smaller than the
recently reported measurements of excess of HXR radiation with respect to thermal emission. For the case of Coma cluster
our predictions, as illustrated in fig. 8, fall short by a factor of several (Fusco-Femiano et al. 1999; Rephaeli et al. 1999).
An analogous estimate, obtained for A2256 after appropriate rescaling, indicates a similar discrepancy by a factor ∼ 30
(Fusco-Femiano et al. 2000), although the upper limits on A3667 are respected by our predictions (Fusco-Femiano et al.
2001). The above discrepancies between prediction and reported detections could be readily improved by assuming that the
electron acceleration efficiency at shocks is larger by an order of magnitude with respect to what assumed here (see also
Loeb & Waxman 2000). Interestingly, since these electrons are located at the cluster outskirts and need not be responsible
for the radio halo emission, their hard X-ray emission would not constraint the ICM magnetic field detected through Faraday
rotation measures. However, as it is apparent from fig. 8, this is at odds with EGRET experimental upper limits on the γ-ray
flux above 100 MeV (Sreekumar et al. 1996; Reimer 1999; Reimer et al. 2003), which allow for the acceleration efficiency
adopted here to be increased by at most a factor ∼ 2 (Miniati 2002b). This constraint, however, holds true only as long
as: (1) CR e− are accelerated above momenta ∼ 100 GeV/c and (2) the spectrum of the accelerated particles is not much
steeper than computed here. These conditions, though, seem to be easily fulfilled for the case of accretion shocks. Thus, in
this respect, if all of the HXR excess emission is diffuse in nature and not associated with individual sources, turbulent and/
or second order Fermi acceleration models may provide a more natural explanation because a high momentum cutoff in this
range arises more naturally in those models (provided of course enough acceleration efficiency).
One of the most awaited experiments is related to the measurement of the γ-ray flux at and above 100 MeV. This is
important in order to convalidate or rule out secondary e± models for radio halos in galaxy clusters (Dennison 1980; Vestrand
1982; Dolag & Ensslin 2000; Blasi & Colafrancesco 1999; Miniati et al. 2001a) and in order to estimate the non-thermal CR
pressure there (Miniati et al. 2001b). In this perspective the above authors estimated for nearby clusters the γ-ray flux
produced from the decay of neutral pions produced in p-p inelastic collisions. As pointed out in Miniati (2002b), however,
γ-ray from IC emission can also be substantial and at the same level as that from π0-decay . Therefore, once again for a
correct interpretation of the measurements, separation of these two components will be required. Estimating the γ-ray flux
from IC emission due to shock accelerated CR electrons will be instrumental for addressing another issue of great interest.
That is, the contribution of this emission mechanism to the cosmic γ-ray background (Loeb & Waxman 2000; Miniati 2002b;
Keshet et al. 2003), for which so far the only constraint is provided by an upper limit based on EGRET experiments (Miniati
2002b).
Photon energies of order ∼ 10 − 100 GeV to 1 − 10 TeV, will be investigated by Cherenkov telescopes. Such measurements
will be complementary to that carried out at lower energies. Because the radiated energy spectrum is directly connected to
the distribution function of the emitting particles, measuring the flux at different energies will provide information about the
acceleration mechanism. In particular, the observed spectrum could differ from our predictions if the accretion shocks were
modified by CR pressure. That in fact would cause the radiation spectrum to soften at low energies and to become harder
toward higher energies, up to an energy cut-off (e.g., Malkov & Drury 2001; Berezhko & Ellison 1999; Kang & Jones 2002).
Given the different environmental conditions with respect to supernova remnants, these are interesting issues for a broad
exploration of shock acceleration physics. Independent of the shock dynamics, measurements in this photon energy range
should help determining the maximum momentum of the accelerated CR electrons and, perhaps, even of CR ions in case
these maxima produce spectral cutoff before attenuation by γγ absorption becomes important. For the electronic component
this is not excluded and could allow a clearer specification of the energy range in which they can contribute to the cosmic
γ-ray background.
5 SUMMARY & CONCLUSIONS
We have explored the spatial and spectral properties of non-thermal emission at γ-ray energies. For the purpose we carried out
a simulation of structure formation including the evolution of baryonic gas, dark matter, CR ions and electrons accelerated
at cosmic shocks as well as secondary e± generated in inelastic p-p collisions. We estimated the radiation flux between 10 keV
and 10 TeV from CRs in collapsed structures due to IC emission, π0-decay and NTB and made specific predictions for the case
of Coma cluster. We have assessed the importance of distinguishing among the contribution from different CR components
for a correct interpretation of future experimental results and we have outlined a strategy to do so. Our conclusions are
summarized as follows:
• Two main regions for production of non-thermal radiation in clusters/groups of galaxies were identified: the core also
bright in thermal X-ray and the outskirts region where accretion shocks occur.
• The chief radiation mechanism at all γ-ray energies in the outskirts region is IC emission from shock accelerated CR
electrons, provided that a fraction of a percent of the shock ram pressure is converted into CR electrons. A clear detection
of this component and of its spatial distribution will allow us direct probing of cosmic accretion shocks. Such evidence would
corroborate recent findings about extra-galactic radio emission from large scale shocks (Bagchi et al. 2002).
• In the cluster core, γ-ray emission above 100 MeV is dominated by π0-decay mechanism. At lower energies, IC emission
from secondary e± takes over. However, IC emission from shock accelerated electrons projected onto the cluster core will not
be negligible in general.
• Separating the aforementioned emission components is important for a correct interpretation of the experimental data.
This can be achieved in principle by measuring the spatial distribution of the detected emission.
• Measuring the non-thermal spectrum will provide us with knowledge regarding: the injection efficiency as well as the
ram pressure to CR pressure conversion efficiency for both electrons and ions; the energy range in which IC emission from
CR electrons accelerated at accretion shocks can contribute to the CGB; the CR content in galaxy clusters; the dynamical
conditions (CR modified or not) of the accretion shocks.
15
ACKNOWLEDGMENTS
This work was carried out at the Max-Planck-Institut fur Astrophysik under the auspices of the European Commission for
the 'Physics of the Intergalactic Medium'. I am grateful to E. Churazov, T. Ensslin, M. Gilfanov, F. Aharonian and I. Susumu
for useful discussion. The computational work was carried out at the the Rechenzentrum in Garching operated by the Institut
fur Plasma Physics and the Max-Planck Gesellschaft.
REFERENCES
Allen, G. E., Petre, R., & Gotthelf, E. V. 2001, ApJ, 558, 739
Arnaud, M., Aghanim, N., Gastaud, R., Neumann, D. M., Lumb, D., Briel, U., Altieri, B., Ghizzardi, S., Mittaz, J., Sasseen,
T. P., & Vestrand, W. T. 2001, A&A, 365, L67
Atoyan, A. M. & Volk, H. J. 2000, ApJ, 535, 45
Bagchi, J., Ensslin, T., Miniati, F., Stalin, C. S., Singh, M., Raychaudhury, S., & Humeshkar, N. B. 2002, NewA, 7, 249
Berezhko, E. G. & Ellison, D. C. 1999, ApJ, 526, 385
Berezinsky, V. S., Blasi, P., & Ptuskin, V. S. 1997, ApJ, 487, 529
Berrington, R. C. & Dermer, C. D. 2002, ApJ
Blasi, P. 1999, ApJ, 525, 603
Blasi, P. & Colafrancesco, S. 1999, Astropart. Phys. , 12, 169
Brunetti, G., Setti, G., Feretti, L., & Giovannini, G. 2001, MNRAS, 320, 365
Colafrancesco, S. & Blasi, P. 1998, Astropart. Phys. , 9, 227
Coppi, P. S. & Aharonian, F. A. 1999, Astropart. Phys. , 11, 35
Deiss, B. M., Reich, W., Lesch, H., & Wielebinski, R. 1997, A&A, 321, 55
Dennison, B. 1980, ApJ, 239, L93
Dolag, K. & Ensslin, T. 2000, A&A, 362, 151
Dwek, E. & Arendt, R. G. 1998, ApJ, 508, L9
Ellison, D. C., Berezhko, E. G., & Baring, M. G. 2000, ApJ, 540, 292
Ellison, D. C. & Eichler, D. 1984, ApJ, 286, 691
Ensslin, T. A. & Biermann, P. L. 1998, A&A, 330, 90
Ensslin, T. A., Biermann, P. L., Kronberg, P. P., & Wu, X.-P. 1997, ApJ, 477, 560
Freedman, W. L. 2000, Phys. Rep., 333, 13
Fusco-Femiano, R., Dal Fiume, D., De Grandi, S., Feretti, L., Giovannini, G., Grandi, P., Malizia, A., Matt, G., & Molendi,
S. 2000, ApJ, 534, L7
Fusco-Femiano, R., Dal Fiume, D., Feretti, L., Giovannini, G., Grandi, P., Matt, G., Molendi, S., & Santangelo, A. 1999,
ApJ, 513, L21
Fusco-Femiano, R., Dal Fiume, D., Orlandini, M., Brunetti, G., Feretti, L., & Giovannini, G. 2001, ApJ, 552, L97
Jones, T. W., Ryu, D., & Engel, A. 1999, ApJ, 512, 105
Kang, H. & Jones, T. W. 1995, ApJ, 447, 994
-- . 2002, JKAS, 35, 159
Keshet, U., Waxman, E., Loeb, A., Springel, V., & Hernquist, L. 2003, ApJ, in press, (astro-ph/0202318)
Kim, K.-T., Kronberg, P. P., Dewdney, P. E., & Landecker, T. L. 1990, ApJ, 355, 29
Kronberg, P. P., Dufton, Q. W., Li, H., & Colgate, S. A. 2001, ApJ, 560, 178
Landau, L. D. & Lifshitz, E. M. 1987, Course of Theoretical Physics, Vol. 6, Fluid Mechanics, 2nd edn. (Oxford: Pergamon
Press)
Loeb, A. & Waxman, E. 2000, Nature, 405, 156
Malkov, M. A. & Drury, L. O. 2001, Rep. Prog. Phys., 64, 429
Mannheim, K. & Schlickeiser, R. 1994, A&A, 286, 983
Meyer, J.-P., Drury, L. O., & Ellison, D. C. 1997, ApJ, 487, 182
Miniati, F. 2000, PhD thesis, University of Minnesota
-- . 2001, Comp. Phys. Comm. , 141, 17
16
F. Miniati
Miniati, F. 2002a, in The Gamma-Ray Universe, ed. A. Goldwurm, D. Neumann, & J. T. T. Van, Moriond Astrophysics
Meeting No. XXIInd, Les Arcs
-- . 2002b, MNRAS, 337, 199
Miniati, F. 2002c, in The Universe Viewed in Gamma-Rays, ed. R. Enomoto, M. Mori, & S. Yanagita (Kashiwa: Universal
Academy Press, Inc.), (preprint astro-ph/0212338)
Miniati, F., Jones, T. W., Kang, H., & Ryu, D. 2001a, ApJ, 562, 233
Miniati, F., Ryu, D., Kang, H., & Jones, T. W. 2001b, ApJ, 559, 59
Miniati, F., Ryu, D., Kang, H., Jones, T. W., Cen, R., & Ostriker, J. 2000, ApJ, 542, 608
Muller, D. & et al. 1995, in Int. Cosmic Ray Conference, Vol. 3, Rome, 13
Muller, D. & Tang, K.-K. 1987, ApJ, 312, 183
Peebles, P. J. E. 1993, Principles of Physical Cosmology (Princeton New Jersey: Princeton University Press)
Quest, K. B. 1988, J. Geoph. Res. , 93, 9649
Reimer, O. 1999, in ICRC, Vol. 4, Int. Cosmic Ray Conference, ed. . B. D. D. Kieda, M. Salamon, Salt Lake City, Utah, 89
Reimer, O., Pohl, M., Sreekumar, P., & Mattox, J. R. 2003, ApJ, in press
Rephaeli, Y., Gruber, D., & Blanco, P. 1999, ApJ, 511, L21
Ryu, D., Ostriker, J. P., Kang, H., & Cen, R. 1993, ApJ, 414, 1
Skilling, J. 1975a, MNRAS, 172, 557
-- . 1975b, MNRAS, 173, 245
Sreekumar, P., Bertsch, D. L., Dingus, B. L., Esposito, J. A., Fichtel, C. E., Fierro, J., Hartman, R. C., Hunter, S. D.,
Kanbach, G., Kniffen, D. A., Lin, Y. C., Mayer-Hasselwander, H. A., Mattox, J. R., Michelson, P. F., von Montigny, C.,
Mukherjee, R., Nolan, P. L., Schneid, E., Thompson, D. J., & Willis, T. D. 1996, ApJ, 464, 628
Vestrand, W. T. 1982, AJ, 87, 1266
Volk, H. J., Aharonian, F. A., & Breitschwerdt, D. 1996, Space Sci. Rev. , 75, 279
White, S. D. M., Navarro, J. F., Evrard, A. E., & Frenk, C. S. 1993, Nature, 366, 429
This figure "fig1.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0303593v1
|
0706.1739 | 1 | 0706 | 2007-06-12T16:11:17 | Speckle noise and dynamic range in coronagraphic images | [
"astro-ph"
] | This paper is concerned with the theoretical properties of high contrast coronagraphic images in the context of exoplanet searches. We derive and analyze the statistical properties of the residual starlight in coronagraphic images, and describe the effect of a coronagraph on the speckle and photon noise. Current observations with coronagraphic instruments have shown that the main limitations to high contrast imaging are due to residual quasi-static speckles. We tackle this problem in this paper, and propose a generalization of our statistical model to include the description of static, quasi-static and fast residual atmospheric speckles. The results provide insight into the effects on the dynamic range of wavefront control, coronagraphy, active speckle reduction, and differential speckle calibration. The study is focused on ground-based imaging with extreme adaptive optics, but the approach is general enough to be applicable to space, with different parameters. | astro-ph | astro-ph |
Speckle noise and dynamic range in coronagraphic images
R´emi Soummer 1
American Museum of Natural History, 79th St. at Central Park West, New York, NY 10024, USA
Accepted for publication in ApJ
[email protected]
and
Andr´e Ferrari, Claude Aime
LUAN, Universit´e de Nice Sophia Antipolis, Parc Valrose, 06108 Nice, France
[email protected], [email protected]
and
Laurent Jolissaint
Herzberg Institute of Astrophysics, 5071 West Saanich Road, Victoria, B.C. V9E 2E7, Canada
[email protected]
ABSTRACT
This paper is concerned with the theoretical properties of high contrast coronagraphic images
in the context of exoplanet searches. We derive and analyze the statistical properties of the
residual starlight in coronagraphic images, and describe the effect of a coronagraph on the speckle
and photon noise. Current observations with coronagraphic instruments have shown that the
main limitations to high contrast imaging are due to residual quasi-static speckles. We tackle
this problem in this paper, and propose a generalization of our statistical model to include the
description of static, quasi-static and fast residual atmospheric speckles. The results provide
insight into the effects on the dynamic range of wavefront control, coronagraphy, active speckle
reduction, and differential speckle calibration. The study is focused on ground-based imaging
with extreme adaptive optics, but the approach is general enough to be applicable to space, with
different parameters.
Subject headings: instrumentation: adaptive optics, instrumentation: high angular resolution
Direct imaging of faint companions or planets around a bright star is a very difficult task, where the
contrast ratio and the angular separation are the observable parameters. The problem consists of detecting
1.
Introduction
1Michelson Fellow, under contract with the Jet Propulsion Laboratory (JPL) funded by NASA through the Michelson
Fellowship Program. JPL is managed for NASA by the California Institute of Technology.
– 2 –
a faint source over a bright and noisy background, mainly due to the diffracted stellar light. High contrast
ratios and small angular separations correspond to the most difficult case. Typically, for extra-solar giant
planets, contrast ratios of about 10−7 are expected in the near infrared (J,H,K bands), based on models
for relatively young ob jects of about 100 Myr (Chabrier & Baraffe 2000; Baraffe et al. 2003; Burrows et al.
2004). According to these models, older ob jects would be an order of magnitude fainter. Terrestrial planets
are much fainter than giant planets, about 3 to 4 orders magnitudes fainter depending on the wavelength
range.
In the case of ground-based observations with Adaptive Optics (AO), the residual uncorrected aberra-
tions produce random intensity fluctuations of the background, which appear as speckles in the field. In direct
non-coronagraphic high quality images, these speckles mainly appear at the position of the diffraction rings
of the star. In Fig.1, we show two high-quality point spread functions (PSF) obtained with the AO system at
Palomar (Troy et al. 2000; Hayward et al. 2001), where the speckles are clearly visible. This phenomenon,
also known as “speckle pinning” (Bloemhof et al. 2001), can be explained using an expansion of the point
spread function (PSF) (Sivaramakrishnan et al. 2002, 2003; Bloemhof 2003a; Perrin et al. 2003; Bloemhof
2004b). An alternative approach using a statistical model (Aime & Soummer 2004b) enables a deeper un-
derstanding of the phenomenon, especially from the statistical point of view, by providing information on
the speckle variance. The effect of a coronagraph on speckle noise is well explained with this approach. In
particular, we show that static aberrations produce residual speckle pinning after a coronagraph, which has
important implications in high contrast imaging. Statistical properties of long exposure AO images were
also studied by Fusco & Conan (2004).
The dynamic range of a coronagraphic image corresponds to the faintest companion that can be detected
at a given position in the field, at the detection limit. It is usually expressed as a magnitude difference or an
intensity ratio, relative to the unocculted central star at some signal to noise ratio (SNR) level. Although
the dynamic range is a two-dimensional map of the detection sensitivity in the field, it is often represented
as a radial profile showing the magnitude difference ∆m as a function of the angular separation r. A radial
profile should be acceptable for most applications, especially when the coronagraph does not have particular
asymmetries (Lyot 1939; Aime et al. 2002; Soummer 2005). Otherwise, a two-dimensional map of sensitivity
is required when the coronagraph shows an asymmetric response (Kuchner & Spergel 2003; Kasdin et al.
2003; Rouan et al. 2000). Understanding, measuring and predicting the dynamic range is still one of the
important issues in this field, with implications for instrumentation (design, observing strategies) and data
reduction and analysis.
During the design phase of future planet finder instruments using Extreme Adaptive Optics (ExAO)
and coronagraphy, for example the Gemini Planet Imager (Macintosh et al. 2004, 2006), or the ESO/VLT
SPHERE (Beuzit et al. 2005; Fusco et al. 2005), it is necessary to anticipate what part of the observable
parameter-space (contrast vs. separation) can be probed, and link it to the actual physical parameter space
(mass vs. semi ma jor axis). Such studies have been carried out by Brown (2004b,a, 2005) in the context of
terrestrial planet searches.
When operating an existing high-contrast instrument, like the Lyot pro ject coronagraph (Oppenheimer et al.
2004), the dynamic range has to be measured to evaluate the performance (Soummer et al. 2006; Hinkley et al.
2006). In the case of a detected ob ject, photometry and astrometry (Digby et al. 2006; Sivaramakrishnan & Oppenheimer
2006; Marois et al. 2006b) are necessary to help determine the ob jects characteristics. Dynamic range com-
putations are also important in the case of non-detection, to determine which part of the parameter space
has been probed by the experiment, and which physical ob jects can be ruled out.
– 3 –
The dynamic range is limited by the intensity fluctuations close to the star. These are due to sev-
eral sources: speckle noise, photon noise, detector noise, background noise etc. Speckle noise is known
to be the main limitation for high contrast imaging, either in direct images (Marois et al. 2003, 2005;
Masciadri et al. 2005; Marois et al. 2006a), or coronagraphic images (Beuzit et al. 1997; Oppenheimer et al.
2001; Boccaletti et al. 2003, 2004; Hinkley et al. 2006). Speckles find their origin in wavefront imperfections
(amplitude and phase errors), whether they correspond to uncorrected atmospheric residual errors (residual
atmospheric speckles) or slow-varying wavefront caused by mechanical or thermal deformations. The main
problem comes from these quasi-static speckles, which can be calibrated either using active pre-focal meth-
ods Malbet et al. (1995); Give’on et al. (2006); Bord´e & Traub (2006) or using post-processing (Marois et al.
2000; Sparks & Ford 2002; Marois et al. 2006a). An alternative approach based on non-redundant masking
has been achieved by Lloyd et al. (2006). Assuming a good enough calibration of these quasi-static speckles,
the physical limit of the system is set by the residual atmospheric aberrations. The noise limitation in
high dynamic range images has been studied by several authors, using numerical simulations (Boccaletti
2004; Cavarroc et al. 2006), or other theoretical or empirical approaches (Angel 1994; Racine et al. 1999;
Perrin et al. 2003; Guyon 2005; Sivaramakrishnan et al. 2005).
In Aime & Soummer (2004b), we modeled the statistics of AO-corrected, direct images, and discussed
the effect in terms of signal to noise ratio on these images qualitatively. In this paper, we examine the effects
a coronagraph has on the properties of residual speckles. A semi-analytical method to compute the dynamic
range can be derived from the statistical properties of the speckle and photon noise. We compare our re-
sults with purely numerical simulations. Results in this paper are presented using an Apodized Pupil Lyot
Coronagraph (APLC) as an example (Aime et al. 2002; Soummer et al. 2003a; Soummer 2005), but are valid
for any other type of mask coronagraph (Rouan et al. 2000; Kuchner & Traub 2002; Soummer et al. 2003c;
Mawet et al. 2005), pure apodizers or shaped pupils (Jacquinot & Roizen-Dossier 1964; Nisenson & Papaliolios
2001; Soummer et al. 2003b; Kasdin et al. 2003; Aime 2005b). Furthermore, the statistical model can be
modified to include both static and quasi-static aberrations, and we discuss the coherent interaction between
residual atmospheric and quasi-static aberrations.
Our theoretical model and results apply for both space and ground-based imaging. However, we illustrate
the results with simulations in the case of ground-based Extreme Adaptive Optics (ExAO) and coronagraphy.
Fig. 1.— Direct images of the bright star HD 137704 (magnitude V=5.47) obtained with the Adaptive
Optics system at the Palomar Hale telescope on June 6th 2004. The core of the image has been saturated
to illustrate the speckles. The bright speckles localized at the position of the diffraction rings are called
pinned speckles. Their statistical properties in direct and coronagraphic images are discussed at length in
this paper.
– 4 –
2. Propagation through a coronagraph in the presence of aberrations
In this section we assume that all the static aberrations, mainly optical quality (polishing for example)
and misalignment errors, can be represented at the entrance pupil of the telescope. The case of out of pupil
aberrations is discussed by Marois et al. (2006c) and does not affect our monochromatic statistical model.
We consider an instrument with an ExAO system and a generic coronagraph which can describe any type of
mask designs (Rouan et al. 2000; Aime et al. 2002; Kuchner & Traub 2002; Soummer et al. 2003c; Soummer
2005; Mawet et al. 2005). The formalism also applies to the shaped pupil approach which corresponds to
the case of direct imaging with apodization (Kasdin et al. 2003). The ExAO and telescope characteristics
are chosen to be consistent with current or future pro jects on eight-meter telescopes and illustrations are
given for an APLC.
These coronagraphs consist of optical filtering in four successive planes denoted 1,2,3,4 hereafter. The
first plane corresponds to the entrance aperture (possibly apodized), the second plane is the focal plane
where a mask is applied (opaque, graded or phase-shifting), the third plane corresponds to a relay pupil
plane where a diaphragm is applied (the Lyot Stop), and finally the fourth plane corresponds to the final
focal plane.
In the pupil plane, we model the wavefront using a static phase aberration ϕs (r), a residual random
atmospheric phase ϕ(r), and amplitude aberrations ρ(r), including the eventual apodization. The complex
amplitude is:
Ψ1(r) = P (r) ρ(r)ej(ϕ(r)+ϕs (r)) ,
(1)
where the function P (r) describes the normalized aperture transmission: R P (r)dr = 1, and r = (x, y ) is the
coordinate vector, used in both pupil and field. Following the notations of Aime & Soummer (2004b), we
can write the wavefront complex amplitude at the entrance pupil as the coherent sum of three terms:
Ψ1 (r) = [A + As (r) + a(r)] P (r),
(2)
where A is a deterministic term corresponding to a perfect plane wave, As (r) is a deterministic complex
term corresponding to the static aberrations, and a(r) is a random term with zero mean (E[a(r)] = 0)
corresponding to the uncorrected part of the wavefront. The probability density function (PDF) of this
complex amplitude is illustrated in Fig.2 without and with AO. Static aberrations are not included in this
figure. A is defined as the mean of the complex amplitude, averaged over the pupil:
A = E[Z Ψ1 (r)P (r)dr].
A2 is therefore the Strehl Ratio of the system (Hardy 1998).
With E[a(r)] = 0, Eq.2 implies E[Ψ1 (r)] = A + As (r). Integrating this equation, we obtain:
E[Z Ψ1 (r)P (r)dr] = A + Z As (r)P (r)dr,
Z As (r)P (r)dr = 0.
Assuming that the phase errors are stationary over the aperture, we obtain:
Z a(r)P (r) ≈ 0.
and therefore:
(6)
(3)
(4)
(5)
– 5 –
Uncorrected
-1. -0.5
0.5
0
1.
SR = 90%
-1. -0.5
0.5
0
1.
SR = 95%
-1. -0.5
0.5
0
1.
1.
0.5
0
-0.5
-1.
1.
1.
1.
1.
0.5
0.5
0
0
-0.5
-0.5
0.5
0.5
0
0
-0.5
-0.5
-1.
-1.
-1.
-1.
1.
0.5
0
-0.5
-1.
-1. -0.5
0
0.5
1.
-1. -0.5
0
0.5
1.
-1. -0.5
0
0.5
1.
Fig. 2.— Probability Density Functions (PDF) of the pupil plane complex amplitude. Left: PDF of un-
corrected atmospheric wavefronts obtained from a von Karman power spectrum model, with an outer scale
L0 = 20m and a seeing ω0 = 0.8 arcsec. The phase excursion is uniform over (0 − 2π) and the thickness of
the annulus corresponds to the amplitude scintillation. Center and Right: PDFs of a AO-corrected wave-
fronts for Strehl Ratio of 90% (center) and 95% (right). These distributions in the complex plane look like
de-centered crescents. The length of the crescent corresponds to the phase excursion, and the thickness to
scintillation, which is assumed to have no effect on the AO system. The effect of the improved AO correction
on the phase excursion is obvious between these two figures.
Note that the difference between Eq.5 and Eq.6 comes from the fact that As (r) is deterministic and a(r) is
random: Although As (r) can be defined with zero-mean over the aperture, each independent realizations of
a(r) do not necessarily have an exactly zero average over the aperture.
The specific case of quasi-static aberrations is not considered in this section and will be treated in Sec.3.2.
The three terms of Eq.2 are illustrated in Fig.3. The length of the vector A is arbitrary in the figure, to
illustrate that the modulus of A is not unity and that the vectors A, As , and a are defined according to the
definitions above. In the first focal plane, a coronagraphic mask is applied at the center of the image of the
ℑ
ϕ(r)
a(r)
As (r)
A
ϕs (r)
ℜ
Fig. 3.— Illustration of the decomposition of the wavefront as the sum of three complex vectors. We consider
a static phase term ϕs and a residual atmospheric phase ϕ. A is the mean wavefront and A2 is therefore
the Strehl Ratio. As corresponds to the static aberrations, and a to the zero-mean error term.
star. Writing the mask transmission as 1 − M (r), allows us to accommodate any type of mask coronagraph,
including Lyot, APLC, Band-Limited, Phase Masks. For example, a classical hard-edged Lyot coronagraph
(or APLC), is described using a top-hat function for M . The complex amplitude of the wave in the focal
– 6 –
plane is given by a scaled Fourier Transform (FT) of this pupil amplitude (Goodman 1996):
Ψ2 (r) = F [Ψ1(r)] (1 − M (r)) ,
where the symbol F denotes the scaled FT. For clarity, we will omit the wavelength-dependent scaling
factors. For the complete chromatic formalism see for example Aime et al. (2002); Soummer et al. (2003a);
Aime (2005a). In the next pupil plane, the complex amplitude before the Lyot stop P ′ (r) is also the sum of
three terms:
(7)
Ψ3 (r) = A Ψc (r) + Ψs (r) + Ψa (r),
where Ψc (r) = P (r) − P (r) ∗ F [M (r)] is the complex amplitude in the lyot stop plane for a perfect wavefront
(Soummer et al. 2003a). The two other terms Ψs (r) and Ψa (r) correspond to the propagation of the terms
As (r) and a(r) respectively. For example:
(8)
Ψa(r) = a(r)P (r) − a(r)P (r) ∗ F [M (r)].
The perfect coronagraph term Ψc (r) and the term Ψa (r) are shown in intensity in Fig.4. The coronagraphs
rejects most of the starlight outside the image of the aperture in the Lyot plane for the perfect part of the
wave, but most of the energy remains inside the aperture for the speckle part. The Lyot stop is applied
(9)
Fig. 4.— Example of propagation though a coronagraph for a monochromatic APLC, corresponding to Eq.8.
Left: intensity in the Lyot stop plane in the perfect case (Ψc (r)2 ) . Most of the light is rejected outside of
the pupil aperture and will be eliminated by the Lyot Stop. Right: intensity of the speckle term Ψa (r)2 ,
assumed alone. Most of the energy remains inside the geometric aperture and will not be eliminated by the
Lyot Stop and appear as residual speckles in the final image. Both images are represented with the same
scale.
in this plane. In the case of an APLC, the Lyot stop is identical to the entrance pupil; in all other cases,
the Lyot stop is undersized. With P (r)P ′ (r) = P ′ (r), and with the notations S (r) = F [a(r) P ′ (r)] and
Ss (r) = F [As (r) P ′ (r)], we obtain the complex amplitude in the final focal plane:
Ψ4 (r) = A Ψd (r)
+Ss (r) − (Ss (r)M (r)) ∗ F [P ′ (r)]
+S (r) − (S (r)M (r)) ∗ F [P ′ (r)],
where Ψd denotes the focal wave amplitude of the coronagraph in the perfect case, following the notations of
Aime et al. (2002). The convolution product (S (r)M (r))∗F [P ′ (r)] in Eq.10 has a negligible effect outside the
mask area. Indeed, the spatial extension of (S (r)M (r)) is limited to the occulting mask area, and F [P ′ (r)]
is a rapidly decreasing function (the Airy amplitude in a perfect case) whose characteristic size is λ/D. The
(10)
– 7 –
result of the convolution does not extend much beyond the mask area, which is illustrated in Fig.5, where
we show an example of the effect of the convolution term using a numerical simulation. This can also be
explained by considering the propagation of phase ripples through a coronagraph and constructing a Bode
diagram (Sivaramakrishnan et al. 2007).
Fig. 5.— Illustration of the terms of Eq.10. Left: modulus of the speckle term alone S (r) = F [a(r) P (r)].
Right: actual term Ψ4 (r) represented in modulus at the same scale. The cross terms of Eq.10 only affect
the central part of the image, inside an area corresponding to the occulting mask size.
Grouping all the deterministic terms together, we introduce
(11)
C (r) = A Ψd (r) + Ss (r) − (Ss (r)M (r)) ∗ F [P ′ (r)],
and we obtain a similar expression to the case without coronagraph where C (r) = A F [P (r)] (Aime & Soummer
2004b):
Ψ4 (r) = C (r) + S (r).
The wave amplitude after a coronagraph appears as a sum of a deterministic term C (r), and a random term
S (r), at each position in the focal plane, outside the mask area. The deterministic term C (r) corresponds to
the focal plane complex amplitude of the coronagraph in the presence of static aberrations. The random term
S (r), associated with the speckles, is identical to the case without coronagraph (Aime & Soummer 2004b):
the coronagraph has a negligible effect on the random part of the wavefront, as illustrated in Fig.4 and Fig.5.
Formally, the effect of the coronagraph is to replace the wave amplitude without coronagraph C (r) by the
coronagraphic amplitude C (r). In the case of pure apodizers (shaped pupils), the direct apodized term C (r)
is used.
(12)
The random term S (r) is non stationary in the field. The profile for S (r) can be computed from a
simulation of the AO system, as we detail in Sec.4.2. Low-order aberrations can also be included in this
description, but usually require a specific study, as for example in Sivaramakrishnan et al. (2005, 2007).
The profile or the two-dimensional map for C (r) can be computed independently, considering a perfect
coronagraph in the presence of deterministic static aberrations (and normalized to the SR). Even in the case
of an ideal coronagraph that cancels all the star light for a perfect wave, the deterministic term C (r) still
contains the terms due to the static aberrations and will contribute to speckle pinning, as discussed below.
– 8 –
3. Statistical properties of direct or coronagraphic images
3.1. Statistical model and properties of speckles
3.1.1. Complex amplitude distribution
(13)
In this section, we discuss the distribution of the complex amplitude in the focal plane. We consider the
case of monochromatic direct images for simplicity. The case of coronagraphic images is formally identical to
the coronagraphic case, according to the approximations described in the previous paragraph (Eq.12). The
focal plane complex amplitude is the Fourier transform of the pupil plane complex amplitude:
Ψ2 (r) = Z P (u)(A + a(u)) e−2ıπu.r du.
The complex amplitude in the focal plane is therefore a sum of the random complex term a(r) weighted by
the Fourier complex phasors. At the center of the image, the Fourier phasors vanish, so a special treatment
for this particular case (and the transition region around it) is necessary (Soummer & Ferrari 2007). Outside
the central point of the image, the distribution of the complex amplitude can be derived using known results
in Signal Processing, based on reasonable assumptions. We assume that the complex amplitude in the pupil
plane can be represented by discrete values (an implicit assumption in any numerical simulation), and that
the correlation of the complex amplitude between two points in the pupil plane decreases with distance
between them. Under these hypotheses, it can be shown that the distribution of the complex amplitude in
the focal plane is asymptotically circular Gaussian (Brillinger 1981). We remind the reader here that if the
real and imaginary parts of a complex number z are Gaussian, its distribution is said to be Gaussian. If
the real and imaginary parts are independent and have same variance, the distribution is said to be circular
Gaussian and denoted z ∼ Nc (0, σ2 ). See Fig.2 of Aime & Soummer (2004b) for an illustration of the focal
plane PDFs. The circularity of the Gaussian distribution is due to the Fourier phasors mixing the complex
amplitudes in the complex plane in the Fourier integral (13), where u varies between −D/2 and D/2. For
positions r in the focal plane such as r > λ/D, the Fourier phase term therefore varies between 0 and 2π and
this circularization occurs. The complex amplitude of the wave in the focal plane Ψ4 (r) follows a circular
Gaussian law, de-centered by the mean of the amplitude C (r) and denoted: Ψ4 (r) ∼ Nc ( C (r), E[S (r)2 ]).
In Fig.6, we give an illustration of the distribution of the complex amplitude in the four successive planes
of the coronagraph. This illustration is based on numerical simulations of a perfect APLC coronagraph and of
an ExAO system. In the first pupil plane, we have a de-centered crescent (see Fig.2). In the first focal plane,
in this example at the top of an Airy ring, C (r) has a high absolute value, and the distribution is Gaussian,
de-centered by this amount. Detailed illustrations of the decentered Gaussian statistics as a function of the
position in the field can be found in Aime & Soummer (2004b). In the following Lyot plane the coronagraph
almost completely removes the perfect part of the wave (see Fig.4), and the resulting distribution is similar
to the initial distribution of the complex amplitude in the pupil, but centered at the origin. Finally in the last
focal plane, without static aberrations, C (r) ≃ 0, as Ψd ≃ 0 and the result is a centered circular Gaussian
distribution.
3.1.2.
Intensity distribution
In this section we derive the Probability Density Function (PDF) of the intensity from that of the wave
complex amplitude. Our problem is formally equivalent to the case of laser speckles added to a coherent
– 9 –
Pupil plane
-1. -0.5
0.5
0
1.
focal plane
-0.1 -0.05 0
0.05 0.1
Lyot Stop plane
-1. -0.5
0.5
0
1.
focal plane
-0.1 -0.05 0
0.05 0.1
1.
0.5
0
-0.5
-1.
1.
0.1
0.5
0.05
0
0
-0.5
-0.05
-1.
-0.1
0.1
1.
0.05
0.5
0
0
-0.05
-0.5
-0.1
-1.
1.
0.1
0.5
0.05
0
0
-0.5
-0.05
-1.
-0.1
0.1
0.05
0
-0.05
-0.1
-1. -0.5
0
0.5
1.
-0.1 -0.05 0
0.05 0.1
-1. -0.5
0
0.5
1.
-0.1 -0.05 0
0.05 0.1
Fig. 6.— Complex probability distributions in the four successive coronagraphic planes (pupil, focal, pupil,
focal), at an arbitrary angular position in the field (r = 2.6λ/D). This simulation corresponds to the case
of an APLC without static aberrations. The distribution in the pupil plane corresponds to a typical ExAO
for an 8-m class telescope delivering 90% SR, and including scintillation effects.
In the focal plane, the
distribution at a given position is a decentered Gaussian.
In the Lyot stop plane after and APLC, the
coronagraph has removed the deterministic part of the entrance pupil wavefront. In the final focal plane,
the distribution is close to circular Gaussian distribution.
background, which has been studied extensively (Goodman 1975, 2006), in particular in the context of
holography. We introduce the two intensity terms:
Ic = C (r)2
Is = E[S (r)2 ].
Note that Ic and Is are both functions of r, and that Ic can describe both the direct or coronagraphic
case, with and without static aberrations. Following Goodman, the joint PDF for the intensity and phase
can be obtained from PDF of the complex amplitude, using the simple cartesian-polar change of variables
(η = √I cos[θ], ξ = √I sin[θ]), where the modulus of the Jacobian of this transformation is kJ k = 1/2 and
integrating the phase θ to find the PDF for the intensity.
(14)
,
(15)
I =
An alternative derivation of the PDF for the intensity is to consider the properties of Gaussian dis-
tributions. As discussed in the previous section, the speckle term S (r) is a circular Gaussian distribution
S (r) ∼ Nc (0, Is ). The instantaneous intensity corresponding to the complex amplitude of Eq.12 is simply:
I = S (r) + C (r)2
= (cid:16)ℜ[ C (r) + S (r)](cid:17)2
+ (cid:16)ℑ[ C (r) + S (r)](cid:17)2
where ℜ and ℑ denote the real and imaginary parts. Using the properties of circular Gaussian distributions,
ℜ[ C (r) + S (r)] and ℑ[ C (r) + S (r)] are independent Gaussian random variables of same variance Is /2. We
can rewrite the intensity with real and imaginary terms of variance unity:
s C (r) + S (r)](cid:17)2(cid:19) =
2 (cid:18)(cid:16)ℜ[p2I −1
s C (r) + S (r)](cid:17)2
Is
+ (cid:16)ℑ[p2I −1
s C (r) + S (r)]] = var[ℑ[p2I −1
where var[ℜ[p2I −1
s C (r) + S (r)]] = 1.
The random variable I follows a de-centered χ2 with two degrees of freedom: χ2
2 (m), with a decentering
s Ic , (Johnson et al. 1995, chap. 29). The probability density function for I is therefore:
parameter m = 2I −1
mv(cid:19) , v > 0,
P (v) = 2−1 e−(m+v)/2f1 (cid:18) 1
4
Is
2
(17)
I ,
(16)
– 10 –
fq (z ) =
1
Γ(q + n)n!
where fq (z ) is the regularized confluent hypergeometric function and 0F1 (; q ; z ) the confluent hypergeometric
function defined as:
z n = 0F1 (; q ; z )
Γ(q)
∞
Xn=0
Finally, the probability density function of the intensity I = Is/2 I is:
0F1 (cid:18); 1;
s (cid:19)
Ic I
I 2
This expression is equivalent 1 to the modified Rician distribution derived by Goodman (1975) and used by
Cagigal & Canales (1998); Canales & Cagigal (1999); Cagigal & Canales (2000); Canales & Cagigal (2001):
Is (cid:19) I0 2√I√Ic
Is ! ,
exp (cid:18)−
I + Ic
where I0 denotes the zero-order modified Bessel function of the first kind. The Rician distribution is illus-
trated in Fig.7. A comparison between the Rician model and simulation data will be presented in Sec.3.1.4.
e− Ic+I
Is
Is
pI (I ) =
1
Is
pI (I ) =
(18)
(19)
(20)
An interesting particular case is when the background C (r) is zero and only the speckle term is present.
Making Ic = 0 in Eq.20 (this happens at the zeros of the perfect PSF or using a perfect coronagraph), the
PDF reduces to:
Is (cid:19) .
exp (cid:18)−
I
This PDF corresponds to the well-known negative exponential density for a fully developed speckle pattern
(e.g. laser speckle pattern) (Goodman 2000). Finally, the distribution at photon counting levels can be ob-
tained performing a Poisson-Mandel transformation of the high flux PDF in Eq.20. An analytical expression
of this PDF has been given in (Aime & Soummer 2004a).
pI (I ) =
1
Is
(21)
The mean and variance of the intensity can be obtained by several ways. A first method (Goodman
1975, 2000), is to express the mean intensity E[I ] and the second order moment of the intensity E[I 2 ] as
a function of C (r) and S (r). The second order moment for the intensity is the fourth order moment for
the complex amplitude: E[I 2 ] = E[(C + S )(C ∗ + S ∗ )]2 ] (omitting the variables r for clarity), which can be
simplified using the properties of Gaussian distributions, with E[S S ∗S S ∗ ] = 2E[S S ∗]E[S S ∗ ] = 2I 2
s we obtain:
c + 4IcIs + 2I 2
E[I 2 ] = I 2
s . A second method is to derive a general analytical expression for the moments of the
Rician distribution. This can be be obtained either from the definition of the moments of Eq.20 (Goodman
1975), or computing the derivatives of the moment generating function (Aime & Soummer 2004a). The
instantaneous intensity in the focal plane (Eq.15) can be written as:
I = C (r)2 + S (r)2 + 2Re[C ∗ (r) S (r)].
Since E[S (r)∗ ] = E[S (r)]∗ = 0 (circular Gaussian distribution), the mean intensity is simply the sum of the
deterministic diffraction pattern with a halo produced by the average of the speckles: Ic + Is or Ic + Is ,
respectively for direct and coronagraphic images. The variance also finds a simple analytical expression, and
(22)
1The Mathematica software (Wolfram 1999) can be used to derive these expressions, and the equivalence between Eq.19 and
Eq.20 can be verified easily using the functions Simplify and FunctionExpand.
– 11 –
P(I)
0.035
0.03
0.025
0.02
0.015
0.01
0.005
P(I)
0.035
0.03
0.025
0.02
0.015
0.01
0.005
P(I)
0.08
0.06
0.04
0.02
0.005
0.01
0.005
0.01
I
I
C(r)
0.2
0.15
0.1
0.05
-0.05
-0.1
2
4
6
r(λ/D)
8
0.005
0.01
I
Fig. 7.— Probability Density Function of the light intensity at 3 different positions in the focal plane,
corresponding to different amplitude C (r) (or intensity levels Ic ). The width of the distribution clearly
increases with an increase in the level of the constant intensity background. This approach provides an
alternative explanation of speckle pinning, where the constant background corresponding to the perfect part
of the wave amplifies speckle fluctuations.
we have:
– 12 –
E[I ] = Is + Ic
σ2
I = I 2
s + 2Is Ic .
(23)
The variance associated with photodetection can be added to this expression to obtain the total variance
P is the variance associated to the poisson statistics: σ2
P , where σ2
I + σ2
σ2 = σ2
P = Ic + Is .
The total variance is therefore:
σ2 = I 2
s + 2Is Ic + Ic + Is .
(24)
In the case of direct images, the term Ic corresponds to the perfect Point Spread Function (PSF) scaled
to the SR. In the case of coronagraphic images, the focal plane intensity is not invariant by translation, and
therefore it is technically not a true PSF. However, we will use the term “coronagraphic PSF” for simplicity
and to follow the general usage in the community. The term Is = E[S (r)2 ] is a function of the radial distance
r, which can describe an actual AO halo. These PSFs and halo structures have been studied analytically
(Moffat 1969; Racine 1996; Racine et al. 1999). It is also possible to determine the halo profile directly from
numerical simulations, and an illustration of Ic and Is is shown in Fig.8. The long exposure PSF profile is
the sum of these two contributions, the halo clearing effect for higher SR (Sivaramakrishnan et al. 2001) is
clearly visible between the two figures. The shape of the halo is due to the spatially filtered wavefront sensor
(Poyneer & Macintosh 2004) used in this simulation. In Fig.9 we show the effect of a coronagraph on the Ic
0
0.2
0.4
r HarcsecL
0.6
0.8
1.
1.2
L
I
H
g
o
L
-5
-5.5
-6
-6.5
-7
-7.5
Ic
Is SR=90%
Is SR=95%
-12.5
-15.
m
D
-17.5
0
5
10
15
r (λ/D)
20
25
30
Fig. 8.— Numerical simulation to illustrate the decomposition of the mean intensity PSF into two compo-
nents Ic and Is for two Strehl Ratios 90% (V=8) and 95% (V=4) for a direct, non-coronagraphic image.
The Is term corresponds to the mean speckle halo and the Ic term corresponds to the perfect PSF, scaled to
the Strehl Ratio, so that the total intensity remains normalized (the difference between the two Ic profiles
is neglected here in log scale). The simulation is made with PAOLA.
term, while the Is term is left unmodified as explained in Sec.2. In this figure we only consider one of the
previous two AO cases. In this example, the coronagraph is good enough to render the constant background
term Ic negligible when compared to the speckle term Is .
– 13 –
0
0.2
0.4
r HarcsecL
0.6
0.8
1.
1.2
cor
Ic
Ic
Is
-10
-15
m
D
-20
-25
-2
-4
-6
-8
-10
L
I
H
g
o
L
0
5
10
15
r (λ/D)
20
25
30
Fig. 9.— Effect of a coronagraph on the Ic term while the Is term is assumed unmodified (Sec.2). We only
consider one case of AO in this figure (SR=95%) and illustrate the Ic term for the direct and coronagraphic
case. The effect of the APLC coronagraph here is to reduce the perfect PSF below the speckle halo. The
corresponding long exposure image is totally dominated by the halo and no residual ringing remains.
3.1.3. Effect of a coronagraph on speckle pinning
In a direct, non coronagraphic image, the term coupling the deterministic C (r) and random parts S (r)
in Eq.22 corresponds to the so-called “speckle pinning”, discussed by several authors (Bloemhof et al. 2001;
Sivaramakrishnan et al. 2002; Bloemhof 2003b; Sivaramakrishnan et al. 2003; Perrin et al. 2003; Bloemhof
2004b), using an expansion of F [P (r)ejϕ(r) ]2 . In this expansion approach, the PSF consist of a sum of
terms, some of which contain a multiplicative factor F [P (r)], corresponding to the diffracted field (the Airy
amplitude for a circular aperture). These terms inherit the zeros of F [P (r)] and contribute all together
to the pinned speckles. Pinned speckles therefore have zero amplitude at the Airy nulls, and non-pinned
speckles remain at these locations. The first order PSF expansion term, denoted by p1 in Perrin et al.
(2003), corresponds to the pinned speckles and the second order (p2halo ) to non-pinned speckles. Higher
order terms contribute to pinned and non-pinned speckles. This expansion approach provides particularly
interesting insight into the spatial properties and symmetries of the speckles for each order of the expansion
(Sivaramakrishnan et al. 2003; Perrin et al. 2003; Bloemhof 2004b). The expansion and our decomposition
are therefore very similar. In Eq.22, our pinning term 2Re[C ∗ (r) S (r)] includes the contribution of all pinning
terms from the infinite expansion. However, our constant term in the decomposition is not 1 and this term
can include the effect of static aberrations, as discussed below.
The analysis of the statistical properties of the speckles enables a deeper understanding of the pinning
phenomenon. As shown in Eq.22 and Eq.23, the pinned speckle term of Eq.22 does not contribute to the
mean intensity, but only contributes to the variance. A numerical simulation is used in Fig.7 to illustrate
the Rician distribution for a direct image, at three different positions in the field: one at the top of an Airy
ring (strong pinning effect), one at a PSF zero (no speckle pinning) and one at an intermediate position.
Speckle pinning can be well illustrated by the analysis of these PDFs, as speckle intensity and fluctuations
are amplified by the term Ic . This can be seen in the PDFs in Fig.7, where the widths increase with Ic .
Depending on the amplitude of the Airy pattern at successive rings, the intensity Ic is alternatively large and
small and the variance of the speckles is amplified accordingly by the coherent part of the wave amplitude
– 14 –
C (r), with corresponding intensity Ic . At the zeroes of the PSF, no amplification occurs and the statistics
is equivalent to that of a fully developed speckle pattern (exponential statistics). It is important to note
that speckles fluctuations are not fully cancelled there (Fig.7), but simply not amplified and their statistics
is that of laser speckles.
The effect of a coronagraph on speckles can be well explained from this statistical modeling. Formally,
as shown in Eq.12, the effect of a coronagraph is to replace the telescope PSF Ic by the on-axis coronagraphic
PSF after a coronagraph Ic . This term includes the effect of static aberrations if they are included in the
model.
• In the perfect case of an ideal coronagraph achieving a total extinction of the star (Rouan et al. 2000;
Aime et al. 2002; Kuchner & Traub 2002; Foo et al. 2005), speckle pinning is fully canceled if there are
no static aberrations in the system.
• In the case of static aberrations in the pupil (due to polishing and alignment errors for example),
these aberrations propagate through the coronagraph, according to the description given in Sec.2, or
in Sivaramakrishnan et al. (2007). The deterministic response of the coronagraph Ic (Eq.11) therefore
includes the effect of these static aberrations. As a result, static aberrations leaking through a coro-
nagraph contribute to speckle pinning, even in the case of a perfect coronagraph. Such effect can be
produced for example by dead actuators on the deformable mirror (Sivaramakrishnan et al. 2006).
Following Aime & Soummer (2004b), we can illustrate the effect of a coronagraph on speckle noise, by
s = I 2
c = 2IcIs + Ic and σ2
breaking down the total variance (Eq.24) into two contributions, σ2
s + Is . The term
σ2
c is the part of the variance that can be removed by a coronagraph by changing Ic , thus affecting speckle
pinning. In Fig.10 and Fig.11 we illustrate these variance contributions for direct and coronagraphic images.
In the case of the direct image, the ringing aspect of the variance corresponds to speckle pinning (where the
variance is amplified). The coronagraphic variance profile is smooth (no amplification of the speckles is seen
at the location of the diffraction rings). Details of the numerical simulation are given in Sec.4.2.
3.1.4. Test of the Rician distribution with numerical simulations
Tests of the Rician statistics on real data using the Lick observatory AO system have been carried out
by Fitzgerald & Graham (2006), showing that the Rician model is consistent with the data. Several compli-
cations exist in real data, so we test the Rician distribution on numerical simulations to determine whether
the model is acceptable in a simple case consistent with the hypothesis used in the physical model, with-
out any additional complicating circumstances or noise. We used PAOLA to generate 10000 AO-corrected
instantaneous phase screens corresponding to an ExAO system, on a 8m telescope (we used a telescope
geometry compatible with Gemini or VLT). The parameters chosen for this simulation include 44 actuators
across the pupil, an integration time and time lag of 0.5ms for a magnitude V = 8 star, and observations in
the H-band. The atmosphere include a typical C 2
n profile for Cerro-Pachon and the seeing is assumed to be
1.4 arcec. The Strehl Ratio of these simulated images is 83%.
In each image, we extract the intensities values along a radius to construct 50 intensity series in the
focal plane at these 50 pixel locations. An example of the first 200 intensity values at an arbitrary location
is given in Fig.12.
For each of these 50 points, we performed a Maximum Likelihood estimation of the parameters Ic and
– 15 –
0
0.2
0.4
r HarcsecL
0.6
0.8
1.
1.2
s2
2
ss
2
sc
2
s
e
c
n
a
i
r
a
v
g
o
l
-4
-5
-6
-7
-8
-9
0
5
10
15
r (λ/D)
20
25
30
Fig. 10.— Total variance for direct imaging (blue solid line) with contributions of the variance terms σs (red
dashed line) and σc which correspond to the pinning contribution (black dotted line). In direct imaging, the
variance budget is totally dominated by the speckle pinning effect, at least close to the axis. This simulation
is made for a SR = 95%.
0
0.2
0.4
r HarcsecL
0.6
0.8
1.
1.2
s2
2
ss
2
sc
-5
-6
-7
-8
-9
-10
-11
2
s
e
c
n
a
i
r
a
v
g
o
l
0
5
10
15
r (λ/D)
20
25
30
Fig. 11.— Total variance for coronagraphic imaging (blue solid line), with contributions of the variance
terms σs (red dashed line) and σc which correspond to the pinning contribution (black dotted line). A
coronagraph can remove the speckle pinning contribution to the variance of the noise. This simulation is
made for a SR = 95% and using and Apodized Pupil Lyot Coronagraph.
– 16 –
y
t
i
s
n
e
t
n
i
0.03
0.02
0.01
0
150
100
50
frame number
200
Fig. 12.— Example of instantaneous PSF simulated with PAOLA, including both phase and amplitude errors
(left) with the intensities for 200 independent frames at an arbitrary position in the focal plane (right). 10000
independent PSFs have been generated at a sampling of 3 times Nyquist, and the values along a radial axis
are stored for this study.
Is , assuming a Rician distribution for the data. The Likelihood has been computed as function of the two
parameters for the un-binned data and maximized using optimization routines of Mathematica. We then
perform the χ2 and Kolomogorov-Smirnoff test statistics on these results. We use ten identical identical bins
for the χ2 test. In Fig.13, we show two examples of binned data with error bars (here due to the Poisson
statistics), superimposed with the Rician distribution for the estimated Is and Ic parameters.
250
200
s
t
n
u
o
C
150
100
50
0
s
t
n
u
o
C
500
400
300
200
100
0
2
4
6
Bin number
8
10
2
4
6
Bin number
8
10
Fig. 13.— Examples of fitted Rician distributions to the binned data, in two cases corresponding to two
different locations in the focal plane: At a top of an Airy ring (left) and at a zero of the perfect PSF where
the distribution is exponential (right). 1000 independent realizations have been used in this simulation.
Our implementation of the Kolmogorov-Smirnoff test is based on a Monte-Carlo estimation of the
distribution.
Indeed, when parameters are estimated from the data (here Ic and Is ), the distribution of
the KS test is not known analytically and must be estimated from Monte-Carlo simulations. For that, we
calculated a distribution of KS test values for 100,000 random drawings of random Rician data where the
parameters have to be estimated from the data. This empirical distribution of the KS test values is used to
generate the empirical right tail values for the test statistic.
The results of the two tests (right tail values) are given in Fig.14, for each of the 50 points along the
radial axis (between 0 and 8.3λ/D). It is interesting to note that for the first two points (on-axis intensity
distribution), both tests conclude that the Rician model is incompatible with the data at the 5% level.
Although the statistical properties at the central point are not directly relevant for evaluating the detection
limits in high contrast images, this question is important in itself as it corresponds to the distribution of Strehl
Ratio. This problem has been studied by Soummer & Ferrari (2007) where it is shown that the distribution
at the center of the image is given by a “reversed” non-central Gamma distribution. This problem was
studied independently by Gladysz et al. (2006); Gladysz (2006); Christou et al. (2006) with similar results.
– 17 –
Outside of the center, most pixels pass the two tests except for about 2 pixel locations, which are rejected,
by one test or the other. This result is consistent with the 5% level we chose, given the total number of
locations (50 pixels) and we conclude that the Rician model is compatible with the simulated data. We
have verified on a few sets of simulations that the location of these pixel failing the test is not relevant and
simply due to the statistics. On the contrary, the central pixel fails the tests systematically. It is interesting
1
0.8
0.6
0.4
0.2
0
0
χ2
Kolmogorov Smirnoff
1
0.8
0.6
0.4
0.2
0
0
5%
10
20
30
40
50
5%
10
20
30
40
50
Fig. 14.— χ2 and KS test results for the 50 locations along a radius in the focal plane. The first point
corresponds to the on-axis case, and the last point is located at 8.3λ/D, with a separation of λ/(6D)
between each point. The solid line shows the 5% rejection threshold. The first two points are rejected by
both tests, which illustrates the non-Rician distribution. In the rest of the field both tests are consistent
with the Rician distribution hypothesis at the 5% rejection level.
to compare the estimated parameters with the actual parameters known from the simulation.
In Fig.15
we compare the actual Ic profile known from the simulation data and compare it to the reconstructed Ic
profile from the estimation of the Rician statistics. Both curves show a very good agreement, even somewhat
surprisingly at the center where the Rician model is wrong. This result is compatible with the satisfactory
results of the test statistics. Fitzgerald & Graham (2006) discussed this type of approach as a possible way
to reconstruct a telescope perfect PSF from the statistics of the PSFs. They note that this procedure might
be difficult to implement with real data, but is at least verified in simulations in this simple case.
Estimated Ic
True Ic
)
c
I
(
g
o
l
0
-5
-10
-15
0
10
20
30
Pixel position
40
50
Fig. 15.— Comparison between the actual perfect term Ic from the simulation without wavefront errors
(red dashed lines), and the Ic values retrieved from the Maximum Likelihood estimation of the parameters
(solid line), assuming the Rician distribution for the intensity. Both curves show a very good agreement,
consistent with the test statistics results. Log intensities are plotted as a function of the radial position in
pixels for easier comparison with other figures.
– 18 –
3.2. Effect of quasi-static aberrations on the statistical properties
Observations at high contrast have shown that the main dynamic range limitations are due to long-lived
quasi-static speckles (Beuzit et al. 1997; Oppenheimer et al. 2001; Marois et al. 2003; Boccaletti et al. 2003,
2004; Soummer et al. 2006). Practical solutions have been proposed to overcome the quasi-static speckles
and will be implemented on the next generation of high contrast imagers (GPI, Sphere) (Wallace et al. 2006;
Marois et al. 2006a) or considered for space pro jects (Bord´e & Traub 2006; Give’on et al. 2006). In ExAO
coronagraphic images, Hinkley et al. (2006) have studied residual speckle lifetimes, and evidenced two types
of speckles, one with lifetimes of a few seconds and the other with lifetimes of a few hundred seconds.
Although there is no reason to believe the universality of these particular values, this problem of quasi-static
speckles is general enough to need further understanding and theoretical insight. Quasi-static speckles are
produced by the slow deformations of the optics (thermal and mechanical) that happen for example as the
telescope slews to follow the star. The typical time scale for these slow variations of aberrations are tens of
minutes.
We can generalize the approach of Sec.2 to include two types of aberrations with different time scales.
We will assume that the error wavefront at the entrance pupil consists of the coherent addition of two
terms: a fast-varying wavefront related to atmospheric residuals, and a quasi-static aberration term. We
assume that quasi-static aberrations can be described as the sum of a deterministic static aberration and a
slowly evolving, zero-mean complex aberration: the static aberrations correspond to the polishing errors and
aberrations produced by the optical design and actual misalignments in the system. Quasi-static aberrations
corresponds to slow zero-mean fluctuations around this static (and therefore deterministic) contribution.
Under these hypothesis, we can add a fourth term to the wavefront decomposition of Eq.2:
Ψ1 (r) = A + As (r) + a1 (r) + a2 (r),
(25)
where, as previously, A corresponds to the perfect wave and As (r) to the static aberrations (deterministic).
The error terms a1 (r) and a2 (r) correspond respectively to the AO-corected and quasi-static aberrations.
We assume that the lifetime of a1 (r) and a2 (r) are respectively τ1 and τ2 > τ1 . We consider the longest
lifetime τ2 as the time unit, and we denote by N the number of fast-speckle realizations during a slow-speckle
lifetime:
N1
τ2
N2
τ1
where N1 and N2 are the number of realizations for each processes during a long exposure of duration ∆t.
Following the results of the previous sections, we can write the field amplitude in the focal plane during N
successive short time intervals τ1 as:
(26)
N =
=
,
Ψ4 (r) = C (r) + S [k]
1 (r) + S2 (r), k = 1, . . . , N .
(27)
As previously, C (r) = C (r) + Ss (r) where C (r) corresponds to the focal plane field (direct of coronagraphic)
in the perfect case without aberrations, and Ss (r) to the static aberration (e.g. polishing errors or actual
optical design aberrations). S [k]
1 (r) corresponds to the AO-corrected atmospheric aberrations and S2 (r) to
the quasi-static aberrations. Outside of the central point in the focal plane, S [k]
1 (r) and S2 (r) are independent
circular Gaussian distributions with zero mean: S [k]
1 (r) ∼ Nc (0, Is1 ) and S2 (r) ∼ Nc (0, Is2 ) with respective
lifetimes τ1 and τ2 . Also, the spatial properties of S [k]
1 (r) and S2 (r) are different as they result from
different aberrations sources, and their corresponding variances Is1 and Is2 can be generated considering the
appropriate spatial power spectra.
– 19 –
During a time unit τ2 corresponding to the lifetime of a slow speckle, the intensity is:
I =
=
(28)
C (r) + S [k]
1 (r) + S2 (r)2
N
Xk=1
N
Xk=1
The expected value of Uk 2 is E[Uk 2 ] = var[Uk ] + E[Uk ]2 , where Uk follows the Gaussian distribution
Uk ∼ Nc ( C (r), Is2 + Is1 ). Therefore, the mean intensity is:
E[I ] = N (Is2 + Is1 + Ic ).
Reminding that Ic = C (r)2 . This expression is consistent with the simple case with one type of speckle
(Eq.23), as it is expressed here for an exposure τ2 = N τ1 .
Uk 2 .
(29)
The variance of the intensity is defined as:
var[I ] =
N
Xk=1
var[Uk 2 ] + Xk 6=l
cov[Uk 2 , Ul 2 ]
= N var[Uk 2 ] + N (N − 1)cov[Uk 2 , Ul 2 ].
The covariance cov[Uk 2 , Ul 2 ] can be easily computed using the expansion of the high order moments of
a complex Gaussian vector as a function of its high order cumulants (McCullagh 1987; Ferrari 2006). The
covariance finds a simple expression:
(30)
l ]2 + E[U ∗
cov[Uk 2 , Ul 2 ] = cov[Uk , U ∗
l ]cov[Ul , U ∗
l ] + E[Uk ]E[U ∗
k ]E[Ul ]cov[Uk , U ∗
k ].
With Uk ∼ Nc ( C (r), Is2 + Is1 ) and k 6= l cov[Uk , U ∗
l ] = Is2 , we obtain immediately the covariance and the
variance (for k=l):
(31)
cov[Uk 2 , Ul 2 ] = I 2
s2 + 2Ic Is2
var[Uk 2 ] = (Is2 + Is1 )2 + 2Ic (Is2 + Is1 )
Finally, the variance of the intensity (Eq.30), for an exposure τ2 , becomes:
s1 + N I 2
var[I ] = N (I 2
s2 + 2Ic (Is1 + N Is2 ) + 2Is1 Is2 )
= N σ2
I ,
where the variance σ2
I for a short exposure τ1 :
s1 + N I 2
I = I 2
σ2
s2 + 2Ic (Is1 + N Is2 ) + 2Is1 Is2
is consistent with the case with one type of speckle (Eq.23), if Is2 = 0.
(32)
(33)
(34)
• In this generalized expression of the variance, the pinning term 2Ic (Is1 + N Is2 ), still exist. However,
it now includes a term corresponding to the pinning of quasi-static speckle. The coefficient N (the
ratio of the speckle lifetimes) can be very large, which makes this pinning particularly efficient: a
quasi-static halo Is2 which is N times lower than Is1 produces the same pinning effect in direct images.
– 20 –
This pinning contribution is still directly affected by a coronagraph, which reduces (or cancels) the Ic
term. In the case of a real, imperfect coronagraph, although no atmospheric pinning might be present,
residual pinning of quasi-static speckles may occur due to the amplification by the factor N if the
quasi-static aberrations are not sufficiently small.
• The cross term 2Is1 Is2 corresponds to speckle pinning of the atmospheric by quasi-static aberrations
If the system is
(coherent amplification of the atmospheric speckles by the quasi-static speckles).
dominated by the quasi-static speckles (Is2 large), this term is negligible compared to the halo N I 2
s2 .
It is possible to specify the performance of a coronagraph to avoid speckle pinning, using this expression for
the variance. Pinning terms are negligible compared to the halo terms if:
2Ic (Is1 + N Is2 ) ≪ I 2
s1 + N I 2
s2 ,
assuming the cross term negligible. In the case of only one type of speckles, or when the system is dominated
by quasi-static speckles, this condition simply becomes Ic ≪ Is (Aime & Soummer 2004b). The necessary
rejection of the coronagraph can be defined as a ratio of the coronagraphic PSF I coro
to the direct PSF
c
I direct
. A similar approach was used by Bloemhof (2004a) to quantify the coronagraph effect on pinned
c
speckles using the expansion approach.
(35)
4. Dynamic range in coronagraphic images
In this paper we adopt a definition of the dynamic range based on an expression of the Signal to Noise
Ratio (S/R), assuming a typical 5σ level detection limit. Based on the knowledge of the noise statistics, more
advanced methods are possible, where the dynamic range can be defined in terms of probability of detection
and false alarm (Michel & Ferrari 2003; Ferrari et al. 2006).
In particular, because of the non-Gaussian
statistics (exponential distribution for a perfect coronagraph), confidence levels have to be carefully defined,
as they do not correspond to the usual values for a Gaussian distribution, C. Marois et al.
(2007, in
preparation). In this section we define analytically the dynamic range in a coronagraphic experiment, using
the variance expressions obtained above. This approach enables a semi-analytical method. We present some
simulations results for the dynamic range.
4.1. Signal to Noise ratio and dynamic range
For simplicity in the calculation of the variance (Sec.3.2), we expressed the intensity for a time unit τ2 as
a sum of short exposures τ1 (Eq.28). In the expression of the variance we derived (Eq.33), the terms Is1 , Is2 ,
Ic correspond implicitly to a number of photons during an exposure τ1 . However, it is convenient in practice
to define normalized intensity terms for a single photon at the entrance aperture and scale them using the
star flux F⋆ . A normalization term τ 2
1 F⋆ is therefore necessary in order to use such normalized intensities.
Recalling that a long exposure consists of N2 exposures of durations τ2 , we obtain the final expression of the
variance for a long exposure ∆t:
1 F 2
speckle = N2 τ 2
σ2
var[I ]
⋆
s1 + N I 2
⋆ (I 2
1 F 2
= N1 τ 2
s2 + 2Ic (Is1 + N Is2 ) + 2Is1 Is2 ).
(36)
– 21 –
The photon noise is obtained by applying the same normalization, multiplying Eq.29 by N2 τ 2
1 :
σ2
photon = (Ic + Is1 + Is2 ) F⋆ ∆t,
(37)
and introducing σ2
p = Ic + Is1 + Is2 for consistency with the case with one type of speckle. The signal
from a companion is Fp f (θ) ∆t, where Fp is the planet flux and f (θ) is a function that describes the off-axis
response of the coronagraph to a companion. f (θ) can be computed independently. For simplicity, we use the
maximum intensity of the off-axis companion for the signal, but other approaches such as matched filtering
can also be used (Soummer et al. 2006). The signal to noise ratio (S/N) is:
.
S/N =
Fp f (θ) ∆t
qN1 τ 2
⋆ σ2
I + F⋆ ∆t σ2
1 F 2
p
This expression for the signal to noise ratio can be immediately converted into an expression for the dynamic
range, assuming a given S/N level. The dynamic range, denoted d is simply the intensity ratio between an
off-axis companion and the star that produces a given S/N level: Fp = d F⋆ . Since ∆t = N1 τ1 , we have:
∆t sσ2
f (θ) r τ1
S/N
I +
Other noise contributions can be easily added in this expression if necessary, but a detailed study of any
specific instrument is beyond the scope of this paper.
σ2
p
τ1F⋆
d =
.
(38)
(39)
4.2. Semi-analytical method based on the statistical model
The dynamic range (Eq.39), can be computed directly from numerical simulations of the terms Is1 , Is2 ,
and Ic . We follow the equations used in the model (Eq.3 to Eq.12) to construct numerical estimates for Ic
and Is (or Is1 , Is2 , if applicable). The term A is obtained from its definition (Eq.3). We consider the average
over the aperture and over a few independent realizations, generated with PAOLA (Jolissaint et al. 2006)
and including both amplitude ρ(r) and phase screens ϕ(r). As described in Sec.2, the term A2 corresponds
to the Strehl Ratio, and is used to generate the Ic profiles: direct and coronagraphic normalized PSFs are
calculated in the perfect case without aberrations, and multiplied by A2 .
The term S (r) is generated following the notation of Eq.10: S (r) = F [a(r) P ′ (r)], where a(r) =
ρ(r) expıϕ(r) −A is the zero-mean complex amplitude at the aperture. The Is term is obtained averag-
ing independent realizations of S (r)2 .
It is easy to verify that the total intensity I = Ic + Is remains
correctly normalized with this procedure.
Finally, radial profiles are generated by averaging these results azimuthally. Satisfactory profiles are
obtained with only a few realizations (typically a few tens), where purely numerical simulations require large
numbers of independent realizations to estimate the same variance. Once the various Ic and Is terms have
been generated, they can be readily combined according to Eq.36 or Eq.38 to produce variance or detection
plots. Multiple instrumental cases can be easily studied this way, combining direct or various coronagraphic
images with different cases of AO is done rapidly using our analytical expressions (for example to study
the effect of stellar magnitude on the dynamic range). This method also enables the study of quasi-static
speckles, which is otherwise computationally expensive.
We compared the variance obtained with the analytical model with the variance obtained from a purely
numerical simulation, computing a large number (∼ 1000) of independent realizations. This process is time-
consuming because of the wavefront propagation through the coronagraph for each phase screen. Excellent
– 22 –
agreement with a full numerical simulation of the variance is obtained with a few realizations (∼ 20), using
the semi-analytical method (Fig.16). We note that the model gives an incorrect estimate of the variance
behind the focal plane occulting mask. This is expected given the approximations described in Eq.11 and
Eq.12. This is not a problem when studying dynamic range, as these expressions are weighted by the off-axis
transmission f (θ) of the coronagraph, and results are only relevant outside the inner working angle of the
instrument.
0
0.2
0.4
r HarcsecL
0.6
0.8
1.
1.2
smodel
2
snum1
2
snum2
2
2
s
e
c
n
a
i
r
a
v
g
o
l
-12
-13
-14
-15
-16
0
5
10
15
r (λ/D)
20
25
30
Fig. 16.— Validation of the results obtained with the semi-analytical model, compared to a purely numerical
experiment. The variance obtained from numerical estimation of Ic and Is based on 20 realizations of phase
and amplitude screens. Solid black line: model, dashed red line: variance for 1000 realizations, dotted blue
line: variance estimated spatially on a single frame.
4.3. Results and comparison with numerical simulations
We present some dynamic range simulation results, using the semi-analytical method based on Eq.39.
We consider an eight-meter telescope, an AO system with 44 actuators across the aperture, a spatially filtered
wavefront sensor (Poyneer & Macintosh 2004), and an open loop frequency of 2.5 kHz. The atmospheric
simulation uses typical seeing values and C 2
n profiles at Mauna Kea. A few cases have been simulated,
with stellar magnitudes ranging from V=4 to V=8. Strehl Ratios are obtained between 87% and 95%.
Atmospheric scintillation is also simulated as amplitude screens with PAOLA, which implements the method
of Roddier (1981). We assumed an instrumental throughput of 25% and an arbitrary exposure time of
∆t = 1000s. The atmospheric lifetime is assumed to be 40ms for observations in the H-band and the stellar
magnitude is V=4. The coronagraph in the simulation is an APLC similar to the one under study for GPI
(Soummer 2005). With a 5σ detection level for this simulation, we illustrate the dynamic range (including
photon and speckle noise only) in Fig.17. In this example, photon and speckle noise are approximately at
the same level inside the AO control region. Outside the control region, the dynamic range is set by the
speckle noise level.
It is interesting to note that the relative position of the speckle and photon contributions do not depend
– 23 –
on exposure time (Eq.39). Indeed, only the speckle lifetime and stellar flux have an effect on the relative
contributions of speckle and photon noise to the dynamic range. This is also true when both fast and quasi-
s and σ2
static speckles are present. In terms of the contributions of σ2
c to the dynamic range as defined in
0
0.2
0.4
r HarcsecL
0.6
0.8
L
I
H
g
o
L
-4
-5
-6
-7
-8
1.
1.2
DR
-10
DR (speckles)
DR (photon)
-12.5
-15
m
D
-17.5
-20
0
5
10
15
r (λ/D)
20
25
30
Fig. 17.— Dynamic range simulation for a 1000s exposure (black solid line), showing the contributions of
the photon noise alone (blue dotted line) and speckle noise alone (red dashed line). In this example, speckle
and photon noise are balanced inside the control region and the outer region is dominated by the speckle
noise.
Aime & Soummer (2004b), the APLC coronagraph is almost perfect in this simulation and the part of the
variance corresponding to speckle pinning, σ2
c , has been completely removed. The case illustrated in Fig.10
and Fig.11 corresponds to the same simulation and shows the effect of the APLC coronagraph canceling the
pinning contribution completely.
Finally, we performed a simulation including both fast and quasi-static aberrations. We considered for
that a simple approach where the power spectrum of the quasi-static aberrations follows a power law f −3 ,
with a lifetime of 600s (10min). This choice is consistent with the assumption that quasi-static aberrations
result from polishing errors, usually well described by such power laws. We considered both photon and
speckle noise, according to Eq.39. In this example, although the quasi-static wavefront error is very small
(10nm RMS), it dominates the error budget and limits the dynamic range inside the control radius of the
AO system. This result depends directly on the power spectrum chosen to generate these aberrations. This
could explain the results obtained with the Lyot pro ject coronagraph where the dynamic range plots do not
show the expected halo-clearing region within the control radius of the AO system Hinkley et al. (2006). A
slightly shallower power spectrum, or a higher level of quasi-static aberrations would easily reproduce the
observed dynamic range by filling the cleared region completely. This effect could be due mainly to the
presence of broken actuators (Sivaramakrishnan et al. 2006) in the AO system.
4.4. Discussion on the effects of speckle reduction techniques
The noise budgets expressed in Eq.36 and Eq.37 enable a general understanding of high contrast imaging
for exoplanet detection. In the case of direct imaging without coronagraph, the main limitation is set by
speckle pinning which appears as an amplification of the variance levels at the position of the maxima of the
– 24 –
0
0.2
0.4
r HarcsecL
0.6
0.8
L
I
H
g
o
L
-4
-5
-6
-7
-8
1.
1.2
-10
DR H1000sL
atmos+AO
-12.5
slow t=600s
-15.
m
D
-17.5
-20.
0
5
10
15
r (λ/D)
20
25
30
Fig. 18.— Dynamic range illustration (black solid line) showing the contribution of fast residual atmospheric
errors (red dashed line) and quasi-static aberrations (blue dotted line). The simulation includes both photon
and speckle noise, and the quasi-static aberrations are assumed to follow a f −3 power law with a 600s
lifetime.
diffracted pattern. In the absence of static aberrations, pinning happens at the top of the Airy rings. In the
presence of static aberrations and even with a perfect coronagraph, residual pinning exists at the position
of the maxima of the coronagraphic PSF corresponding to the propagated static aberrations through the
coronagraph. For example, in the case of low-order static aberrations described by Zernike polynomials,
the location of the maxima of the PSF (direct or coronagraphic) dictates where pinning happens. These
locations are no longer at the top of the perfect Airy pattern in this case. Quasi-static aberrations also add
to the problem by creating additional pinning terms (Eq.34). This phenomenon was observed at the AEOS
telescope with the Lyot pro ject coronagraph, where pinned speckles have been associated with dead actuators
on the deformable mirror (Sivaramakrishnan et al. 2006), or AO waffle mode (Makidon et al. 2005).
In reality, coronagraphs are not perfect and suffer from various sources of errors, including chromaticity,
manufacturing imperfections, alignment errors, etc. The cancellation of speckle pinning in the final error
budget provides a constraint on the coronagraph performance. This can be achieved based on the comparison
between the terms Ic and Is described above.
With current state of the art AO systems and coronagraphs alone, it is clear that the residual noise
terms (even if the pinning term is well controlled) would not enable the detection of giant planets with typical
contrasts of 10−7 at 0.5 arcsec, or Earth-like planets with contrast of 10−10 at 0.1 arcsec. Additional speckle
reduction techniques have been developed (Bord´e & Traub 2006; Give’on et al. 2006) to improve the contrast
performance further in combination with wavefront control and coronagraphs. These techniques affect the
term Is by modifying the pupil aberrations to create a dark zone in the search region of the field. In the case
of ground-based high contrast imagers, for example with the GPI speckle calibration system (Wallace et al.
2006), the speckle reduction will only affect the Is2 term, as the sensing involved in the technique is slow.
Some speckle reduction techniques may also affect the residual speckle pinning in some region of the field,
as they create a static wavefront aberration which creates a particularly asymmetric PSF presenting a dark
zone (Codona et al. 2006; Serabyn et al. 2006).
The combination of coronagraphy and speckle nulling will therefore remove most of the terms contribut-
– 25 –
ing to the noise variance σ2
I . The limit for the association of coronagraphy and speckle nulling is set by the
level of residual fast atmospheric aberrations. Also, by decreasing the term Is2 , the speckle reduction system
puts a stronger constraint on the coronagraph, whose performance must suppress pinned speckles to a level
below other speckles in the dark region produced by speckle nulling.
Finally, the remaining parts of the speckle noise which have not been removed by a coronagraph or by
the speckle reduction system, can be removed further using post-processing differential methods, combining
multi-wavelength information from dual-band imaging, or an integral field spectrograph Marois et al. (2000);
Sparks & Ford (2002); Marois et al. (2005), or using differential rotation Marois et al. (2006a). These dif-
ferential methods affect all the residual speckle variance contributions (pinning and halo) in the same way.
The combination of these three stages (coronagraph, speckle reduction and speckle processing) is necessary
to reach the necessary dynamic range regime for exoplanet imaging. This approach has been chosen for the
design of GPI (Macintosh et al. 2006), and have been implemented in the laboratory by Trauger & Traub
(2007).
5. Conclusion
This paper develops theoretical insight in high contrast imaging, using a statistical model to understand
the performance limitations of coronagraphic experiments. We discuss the application of the statistical model
(Aime & Soummer 2004b) to the case of coronagraphic images, and show that the statistical model is valid
outside of the focal plane mask occulting area. We show some limitations of the statistical model at the
central point of the image, and a modification of the model is proposed by Soummer & Ferrari (2007) to
solve this problem.
It has been confirmed by several observations that the main limitation for high contrast imaging is
due to the presence of quasi-static speckles. These do not average out over time, and are difficult to
calibrate. We presented the first theoretical attempt to understand the effect of quasi-static speckles and their
interaction with other aberrations (static and residual atmospheric). We obtained an analytical expression
for the variance of the intensity, which includes speckle and photon noise in the presence of static, quasi-
static and rapidly-varying aberrations. This result enables the use of a semi-analytical method, which has
been compared successfully with purely numerical simulations. This semi-analytical method enables fast
simulations and the possibility to compare and combine easily various types of parameters for instrument
design studies. It also enables a breakdown of the noise into specific contributions (pinning, halo, speckle,
photon). This method can be used to model real observations and extract information on relative noise
contributions from various sources (residual atmospheric noise, quasi-static noise).
The model also provides an understanding of the speckle pinning effect in high Strehl images, where
bright speckles appear at the location of the diffraction rings. This effect, explained by a coherent ampli-
fication of the speckle noise, and can be cancelled by a coronagraph. The theory provides insight on the
required level of performance for the coronagraph to effectively cancel speckle pinning. In the presence of
quasi-static aberrations, we showed that pinning also exists between the perfect part of the wave and quasi-
static aberrations, but is weighted by the ratio of the lifetimes (which can be a large number of the order of
a hundred or more). This puts strong requirements on the level of static and quasi-static aberrations and
may explain the results currently obtained with the Lyot pro ject coronagraph, the first instrument of its
kind to use an extreme AO system. Detail modeling of the Lyot pro ject data, and specific simulations for
the future generation of high contrast imager (GPI), will be done in the future.
– 26 –
This model was developed for the monochromatic case. Generalization to wide band is not straight-
forward, as the speckle noise is highly correlated with wavelength, which makes this generalization difficult,
but can be used for multi-wavelength speckle subtractions. Also, a generalization of the model to the case of
subtraction residuals in multi-wavelength imaging would be particularly interesting and is a topic for future
study.
Although this presentation is focused on the ground-based case, the same formalism can be directly
used for space-based observations, with similar results, but different orders of magnitudes (in particular for
the speckle lifetimes). An identical expression of the variance of the noise was established independently by
Shaklan et al. (2005) for the TPF error budget.
Once a coronagraph has removed the speckle pinning contribution, the dynamic range can be improved
further by use of speckle reduction techniques (speckle nulling) or speckle subtraction (ADI, multi-wavelength
or polarization). The effect of speckle nulling on the noise budget is now well understood from the expression
of the variance. The model can be used to draw the requirements on the relative performance of the
coronagraph and speckle reduction system. This theory brings an understanding of the effects on the dynamic
range of the various elements in a high contrast instrument, including wavefront control, coronagraphy,
speckle reduction and differential calibration techniques. All these stages are crucial and necessary to reach
high contrast imaging, and they are going to be implemented in the next generation of high contrast imagers,
currently in development.
R´emi Soummer is supported by a Michelson Postdoctoral Fellowship, under contract to the Jet Propul-
sion Laboratory (JPL) funded by NASA. JPL is managed for NASA by the California Institute of Technology.
This work is based upon work partially supported by the National Science Fundation under Grant number
No. AST-0215793 and AST-0334916 and has also been supported in part by the National Science Founda-
tion Science and Technology Center for Adaptive Optics, managed by the University of California at Santa
Cruz under cooperative agreement AST 98-76783. The authors would like to thank Julian Christou, Szymon
Gladysz, James P. Lloyd, Louis Lyons, Bruce Macintosh, Russell Makidon, Christian Marois and Anand
Sivaramakrishnan for interesting discussions.
Aime, C. 2005a, PASP, 117, 1012
—. 2005b, A&A, 434, 785
REFERENCES
Aime, C. & Soummer, R. 2004a, in EAS Publications Series, ed. C. Aime & R. Soummer, 89–101
Aime, C. & Soummer, R. 2004b, ApJ, 612, L85
Aime, C., Soummer, R., & Ferrari, A. 2002, A&A, 389, 334
Angel, J. R. P. 1994, Nature, 368, 203
Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, A&A, 402, 701
Beuzit, J.-L., Mouillet, D., Dohlen, K., & Puget, P. 2005, in SF2A-2005: Semaine de l’Astrophysique
Francaise, ed. F. Casoli, T. Contini, J. M. Hameury, & L. Pagani, 215–+
Beuzit, J.-L., Mouillet, D., Lagrange, A.-M., & Paufique, J. 1997, A&AS, 125, 175
– 27 –
Bloemhof, E. E. 2003a, ApJL, 582, L59
—. 2003b, ApJL, 582, L59
—. 2004a, ApJ, 610, L69
—. 2004b, Optics Letters, 29, 2333
Bloemhof, E. E., Dekany, R. G., Troy, M., & Oppenheimer, B. R. 2001, ApJ, 558, L71
Boccaletti, A. 2004, in EAS Publications Series, ed. C. Aime & R. Soummer, 165–176
Boccaletti, A., Chauvin, G., Lagrange, A.-M., & Marchis, F. 2003, A&A, 410, 283
Boccaletti, A., Riaud, P., Baudoz, P., Baudrand, J., Rouan, D., Gratadour, D., Lacombe, F., & Lagrange,
A.-M. 2004, PASP, 116, 1061
Bord´e, P. J. & Traub, W. A. 2006, ApJ, 638, 488
Brillinger, D. 1981, Time Series : Data Analysis and Theory (McGraw-Hill)
Brown, R. A. 2004a, ApJ, 610, 1079
—. 2004b, ApJ, 607, 1003
—. 2005, ApJ, 624, 1010
Burrows, A., Sudarsky, D., & Hubeny, I. 2004, ApJ, 609, 407
Cagigal, M. P. & Canales, V. F. 1998, Optics Letters, 23, 1072
—. 2000, JOSA, 17, 1312
Canales, V. F. & Cagigal, M. P. 1999, AO, 38, 766
—. 2001, Optics Letters, 26, 737
Cavarroc, C., Boccaletti, A., Baudoz, P., Fusco, T., & Rouan, D. 2006, A&A, 447, 397
Chabrier, G. & Baraffe, I. 2000, ARA&A, 38, 337
Christou, J. C., Gladysz, S., Redfern, M., Bradford, L. W., & Roberts, L. C. J. 2006, in Proc AMOS technical
conference, Maui, Hawaii, 10-14 September 2006, 528–537
Codona, J. L., Kenworthy, M. A., Hinz, P. M., Angel, J. R. P., & Woolf, N. J. 2006, in Ground-based and
Airborne Instrumentation for Astronomy. Edited by McLean, Ian S.; Iye, Masanori. Proceedings of
the SPIE, Volume 6269, pp. (2006).
Digby, A. P., Hinkley, S., Oppenheimer, B. R., Sivaramakrishnan, A., Lloyd, J. P., Perrin, M. D., Roberts,
L. C. J., Soummer, R., Brenner, D., Makidon, R. B., Shara, M., Kuhn, J., Graham, J. R., Kalas,
P. G., & Newburgh, L. 2006, Submited to ApJ
Ferrari, A. 2006, in EAS Publications Series, 85–101
– 28 –
Ferrari, A., Carbillet, M., Serradel, E., Aime, C., & Soummer, R. 2006, in Proceedings IAU Colloquium No.
200, ed. C. Aime & J. Vakili, F. Darcourt
Fitzgerald, M. P. & Graham, J. R. 2006, ApJ, 637, 541
Foo, G., Palacios, D. M., & Swartzlander, Jr., G. A. 2005, Optics Letters, 30, 3308
Fusco, T. & Conan, J.-M. 2004, Journal of the Optical Society of America A, 21, 1277
Fusco, T., Rousset, G., Mouillet, D., & Beuzit, J.-L. 2005, in SF2A-2005: Semaine de l’Astrophysique
Francaise, ed. F. Casoli, T. Contini, J. M. Hameury, & L. Pagani, 219–+
Give’on, A., Kasdin, N. J., Vanderbei, R. J., & Avitzour, Y. 2006, JOSA A, 23
Gladysz, S. 2006, PhD thesis, National University of Ireland, Galway
Gladysz, S., Christou, J. C., & Redfern, M. 2006, in Advances in Adaptive Optics II. Edited by Ellerbroek,
Brent L.; Bonaccini Calia, Domenico. Proceedings of the SPIE, Volume 6272, pp. (2006).
Goodman, J. 1975, in topics in applied physics : laser speckle and related phenomena, Dainty Ed. (springer
verlag berlin)
Goodman, J. 1996, Introduction to Fourier Optics (Mac Graw Hill)
Goodman, J. W. 2000, Statistical Optics (Wiley Classics Library), wiley classics library ed. edn., Wiley
classics library (New York: Wiley-Interscience), 576
—. 2006, Speckle Phenomena in Optics (Englewood, Colo.: Roberts and Company Publishers), 384
Guyon, O. 2005, ApJ, 629, 592
Hardy, J. W. 1998, Adaptive Optics for Astronomical Telescopes (Oxford University Press, USA), 448
Hayward, T. L., Brandl, B., Pirger, B., Blacken, C., Gull, G. E., Schoenwald, J., & Houck, J. R. 2001, PASP,
113, 105
Hinkley, S., Oppenheimer, B. R., Soummer, R., Sivaramakrishnan, A., Roberts, L. C. J., Kuhn, J., Makidon,
R. B., Perrin, M. D., Lloyd, J. P., Kratter, K., & Brenner, D. 2006, Submited to ApJ
Jacquinot, P. & Roizen-Dossier, B. 1964, Progress in Optics, Vol. 3 (Wolf, E.)
Johnson, N., Kotz, S., & Balakrishnan, N. 1995, Continuous Univariate Distributions, Vol. 2 (John Wiley &
Sons Inc.)
Jolissaint, L., V´eran, J.-P., & Conan, R. 2006, Journal of the Optical Society of America A, 23, 382
Kasdin, N. J., Vanderbei, R. J., Spergel, D. N., & Littman, M. G. 2003, ApJ, 582, 1147
Kuchner, M. J. & Spergel, D. N. 2003, ApJ, 594, 617
Kuchner, M. J. & Traub, W. A. 2002, ApJ, 570, 900
Lloyd, J. P., Martinache, F., Ireland, M. J., Monnier, J. D., Pravdo, S. H., Shaklan, S. B., & Tuthill, P. G.
2006, ApJ, 650, L131
– 29 –
Lyot, B. 1939, MNRAS, 99, 580
Macintosh, B., Graham, J., Palmer, D., Doyon, R., Gavel, D., Larkin, J., Oppenheimer, B., Saddlemyer,
L., Wallace, J. K., Bauman, B., Evans, J., Erikson, D., Morzinski, K., Phillion, D., Poyneer, L.,
Sivaramakrishnan, A., Soummer, R., Thibault, S., & Veran, J.-P. 2006, in Advances in Adaptive
Optics II. Edited by Ellerbroek, Brent L.; Bonaccini Calia, Domenico. Proceedings of the SPIE,
Volume 6272, pp. (2006).
Macintosh, B. A., Bauman, B., Wilhelmsen Evans, J., Graham, J. R., Lockwood, C., Poyneer, L., Dillon,
D., Gavel, D. T., Green, J. J., Lloyd, J. P., Makidon, R. B., Olivier, S., Palmer, D., Perrin, M. D.,
Severson, S., Sheinis, A. I., Sivaramakrishnan, A., Sommargren, G., Soummer, R., Troy, M., Wallace,
J. K., & Wishnow, E. 2004, in Advancements in Adaptive Optics. Edited by Domenico B. Calia, Brent
L. Ellerbroek, and Roberto Ragazzoni. Proceedings of the SPIE, Volume 5490, pp. 359-369 (2004).,
ed. D. Bonaccini Calia, B. L. Ellerbroek, & R. Ragazzoni, 359–369
Makidon, R. B., Sivaramakrishnan, A., Perrin, M. D., Roberts, L. C., Oppenheimer, B. R., Soummer, R.,
& Graham, J. R. 2005, PASP, 117, 831
Malbet, F., Yu, J. W., & Shao, M. 1995, PASP, 107, 386
Marois, C., Doyon, R. ., Racine, R. ., & Nadeau, D. 2000, PASP, 112, 91
Marois, C., Doyon, R., Nadeau, D., Racine, R., Riopel, M., Vall´ee, P., & Lafreni`ere, D. 2005, PASP, 117,
745
Marois, C., Doyon, R., Nadeau, D., Racine, R., & Walker, G. A. H. 2003, in EAS Publications Series, ed.
C. Aime & R. Soummer, 233–243
Marois, C., Lafreni`ere, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006a, ApJ, 641, 556
Marois, C., Lafreni`ere, D., Macintosh, B., & Doyon, R. 2006b, ApJ, 647, 612
Marois, C., Phillion, D. W., & Macintosh, B. 2006c, in Ground-based and Airborne Instrumentation for
Astronomy. Edited by McLean, Ian S.; Iye, Masanori. Proceedings of the SPIE, Volume 6269, pp.
(2006).
Masciadri, E., Mundt, R., Henning, T., Alvarez, C., & Barrado y Navascu´es, D. 2005, ApJ, 625, 1004
Mawet, D., Riaud, P., Absil, O., & Surdej, J. 2005, ApJ, 633, 1191
McCullagh, P. 1987, Tensor methods in statistics, Monographs on statistics and applied probability (Wiley,
New York)
Michel, O. & Ferrari, A. 2003, in EAS Publications Series, ed. C. Aime & R. Soummer, 129–146
Moffat, A. F. J. 1969, A&A, 3, 455
Nisenson, P. & Papaliolios, C. 2001, ApJ Letters, 549
Oppenheimer, B. R., Digby, A. P., Newburgh, L., Brenner, D., Shara, M., Mey, J., Mandeville, C., Makidon,
R. B., Sivaramakrishnan, A., Soummer, R., Graham, J. R., Kalas, P., Perrin, M. D., Roberts, L. C.,
Kuhn, J. R., Whitman, K., & Lloyd, J. P. 2004, in Advancements in Adaptive Optics. Edited by
Domenico B. Calia, Brent L. Ellerbroek, and Roberto Ragazzoni. Proceedings of the SPIE, Volume
5490, pp. 433-442 (2004)., ed. D. Bonaccini Calia, B. L. Ellerbroek, & R. Ragazzoni, 433–442
– 30 –
Oppenheimer, B. R., Golimowski, D. A., Kulkarni, S. R., Matthews, K., Naka jima, T., Creech-Eakman, M.,
& Durrance, S. T. 2001, AJ, 121, 2189
Perrin, M. D., Sivaramakrishnan, A., Makidon, R., Oppenheimer, B. R., & Graham, J. R. 2003, ApJ, 596,
702
Poyneer, L. A. & Macintosh, B. 2004, Journal of the Optical Society of America A, vol. 21, Issue 5, pp.810-
819, 21, 810
Racine, R. 1996, PASP, 108, 699
Racine, R., Walker, G. A. H., Nadeau, D., Doyon, R., & Marois, C. 1999, PASP, 111, 587
Roddier, F. 1981, Prog. Optics, Volume 19, p. 281-376, 19, 281
Rouan, D., Riaud, P., Boccaletti, A., Cl´enet, Y., & Labeyrie, A. 2000, PASP, 112, 1479
Serabyn, E., Wallace, J. K., Troy, M., Mennesson, B., Haguenauer, P., Gappinger, R. O., & Bloemhof, E. E.
2006, in Advances in Adaptive Optics II. Edited by Ellerbroek, Brent L.; Bonaccini Calia, Domenico.
Proceedings of the SPIE, Volume 6272, pp. (2006).
Shaklan, S. B., Marchen, L., Green, J. J., & Lay, O. P. 2005, in Techniques and Instrumentation for Detection
of Exoplanets II. Edited by Coulter, Daniel R. Proceedings of the SPIE, Volume 5905, pp. 110-121
(2005)., ed. D. R. Coulter, 110–121
Sivaramakrishnan, A., Hodge, P. E., Makidon, R. B., Perrin, M. D., Lloyd, J. P., Bloemhof, E. E., &
Oppenheimer, B. R. 2003, in High-Contrast Imaging for Exo-Planet Detection. Edited by Alfred B.
Schultz. Proceedings of the SPIE, Volume 4860, pp. 161-170 (2003)., 161–170
Sivaramakrishnan, A., Koresko, C. D., Makidon, R. B., Berkefeld, T., & Kuchner, M. J. 2001, ApJ, 552, 397
Sivaramakrishnan, A., Lloyd, J. P., Hodge, P. E., & Macintosh, B. A. 2002, ApJL, 581, L59
Sivaramakrishnan, A. & Oppenheimer, B. R. 2006, ApJ, 647, 620
Sivaramakrishnan, A., Oppenheimer, B. R., Perrin, M. D., Roberts, L. C., Makidon, R. B., Soummer, R.,
Digby, A. P., Bradford, L. W., Skinner, M. A., Turner, N. H., & Ten Brummelaar, T. A. 2006, in IAU
Colloq. 200: Direct Imaging of Exoplanets: Science & Techniques, ed. C. Aime & F. Vakili, 613–616
Sivaramakrishnan, A., Soummer, R., Oppenheimer, B., & Pueyo, L. 2007, Submited to ApJ
Sivaramakrishnan, A., Soummer, R., Sivaramakrishnan, A. V., Lloyd, J. P., Oppenheimer, B. R., & Makidon,
R. B. 2005, ApJ, 634, 1416
Soummer, R. 2005, ApJ, 618, L161
Soummer, R., Aime, C., & Falloon, P. E. 2003a, A&A, 397, 1161
Soummer, R., Aime, C., Ferrari, A., & Falloon, P. E. 2003b, in High-Contrast Imaging for Exo-Planet
Detection. Edited by Alfred B. Schultz. Proceedings of the SPIE, Volume 4860, pp. 211-220 (2003).,
ed. A. B. Schultz, 211–220
Soummer, R., Dohlen, K., & Aime, C. 2003c, A&A, 403, 369
– 31 –
Soummer, R. & Ferrari, A. 2007, ApJL, in press
Soummer, R., Oppenheimer, B. R., Hinkley, S., Sivaramakrishnan, A., Makidon, R. B., Oppenheimer, B. R.,
Digby, A. P., Brenner, D., Kuhn, J., Perrin, M. D., Roberts, L. C. J., & Kratter, K. 2006, in EAS
Publications Series, ed. C. Aime & M. Carbillet
Sparks, W. B. & Ford, H. C. 2002, ApJ, 578, 543
Trauger, J. T. & Traub, W. A. 2007, Nature, 446, 771
Troy, M., Dekany, R. G., Brack, G., Oppenheimer, B. R., Bloemhof, E. E., Trinh, T., Dekens, F. G., Shi,
F., Hayward, T. L., & Brandl, B. 2000, in Proc. SPIE Vol. 4007, p. 31-40, Adaptive Optical Systems
Technology, Peter L. Wizinowich; Ed., 31–40
Wallace, J. K., Bartos, R., Rao, S., Samuele, R., & Schmidtlin, E. 2006, in Advances in Adaptive Optics II.
Edited by Ellerbroek, Brent L.; Bonaccini Calia, Domenico. Proceedings of the SPIE, Volume 6272,
pp. (2006).
Wolfram, S. 1999, The Mathematica book, Fourth Edition (Cambridge University Press)
This preprint was prepared with the AAS LATEX macros v5.2.
|
0801.4763 | 2 | 0801 | 2008-02-07T16:47:31 | Gas and Dust Emission at the Outer Edge of Protoplanetary Disks | [
"astro-ph"
] | We investigate the apparent discrepancy between gas and dust outer radii derived from millimeter observations of protoplanetary disks. Using 230 and 345 GHz continuum and CO J=3-2 data from the Submillimeter Array for four nearby disk systems (HD 163296, TW Hydrae, GM Aurigae, and MWC 480), we examine models of circumstellar disk structure and the effects of their treatment of the outer disk edge. We show that for these disks, models described by power laws in surface density and temperature that are truncated at an outer radius are incapable of reproducing both the gas and dust emission simultaneously: the outer radius derived from the dust continuum emission is always significantly smaller than the extent of the molecular gas disk traced by CO emission. However, a simple model motivated by similarity solutions of the time evolution of accretion disks that includes a tapered exponential edge in the surface density distribution (and the same number of free parameters) does much better at reproducing both the gas and dust emission. While this analysis does not rule out the disparate radii implied by the truncated power-law models, a realistic alternative disk model, grounded in the physics of accretion, provides a consistent picture for the extent of both the gas and dust. | astro-ph | astro-ph | Accepted for publication in ApJ: January 29, 2008
Preprint typeset using LATEX style emulateapj v. 08/22/09
8
0
0
2
b
e
F
7
]
h
p
-
o
r
t
s
a
[
2
v
3
6
7
4
.
1
0
8
0
:
v
i
X
r
a
GAS AND DUST EMISSION AT THE OUTER EDGE OF PROTOPLANETARY DISKS
A. M. Hughes 1, D. J. Wilner 1, C. Qi 1, M. R. Hogerheijde 2
Accepted for publication in ApJ: January 29, 2008
ABSTRACT
We investigate the apparent discrepancy between gas and dust outer radii derived from millimeter
observations of protoplanetary disks. Using 230 and 345 GHz continuum and CO J=3-2 data from
the Submillimeter Array for four nearby disk systems (HD 163296, TW Hydrae, GM Aurigae, and
MWC 480), we examine models of circumstellar disk structure and the effects of their treatment of
the outer disk edge. We show that for these disks, models described by power laws in surface density
and temperature that are truncated at an outer radius are incapable of reproducing both the gas and
dust emission simultaneously: the outer radius derived from the dust continuum emission is always
significantly smaller than the extent of the molecular gas disk traced by CO emission. However, a
simple model motivated by similarity solutions of the time evolution of accretion disks that includes a
tapered exponential edge in the surface density distribution (and the same number of free parameters)
does much better at reproducing both the gas and dust emission. While this analysis does not rule
out the disparate radii implied by the truncated power-law models, a realistic alternative disk model,
grounded in the physics of accretion, provides a consistent picture for the extent of both the gas and
dust.
Subject headings: accretion, accretion disks -- circumstellar matter -- planetary systems: protoplan-
etary disks -- stars: pre-main sequence
1. INTRODUCTION
Characterizing the gas and dust distribution in the
disks around young stars is important for understanding
the planet formation process, as these disks provide the
reservoirs of raw material for nascent planetary systems.
A common method of modeling circumstellar disk struc-
ture is to use models described by power laws in surface
density and temperature that are truncated at a particu-
lar outer radius. This prescription has its historical roots
in calculations of the minimum mass solar nebula, which
indicated a surface density profile of Σ ∝ r−3/2 (e.g.
Weidenschilling 1977), as well as theoretical predictions
of a radial power-law dependence of temperature for ac-
creting disks around young stars (Adams & Shu 1986;
Adams et al. 1987). Observationally, the parameteri-
zation of temperature and surface density as power-law
functions of radius began with early spatially unresolved
studies of continuum emission from disks (Beckwith et al.
1990; Beckwith & Sargent 1991). These models have
since been refined and applied to spatially resolved ob-
servations of many disks with success (e.g. Mundy et al.
1993; Dutrey et al. 1994; Lay et al. 1994; Dutrey et al.
1998), and they have proven useful for understanding
the basic global properties of disk structure. Recently,
however, with the advent of high signal-to-noise, multi-
frequency observations of gas and dust in protoplanetary
disks, these models have begun to encounter difficulties,
particularly in the treatment of the outer disk edge.
The extent of the gas and dust distribution in circum-
stellar disks has implications for our understanding of
the planet formation process in our own solar system.
Electronic address: [email protected]
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden
Street, Cambridge, MA 02138
2 Leiden Observatory, Leiden University, P.O. Box 9513, 2300
RA, Leiden, The Netherlands
There is some evidence for a sharp decrease in the sur-
face density of Kuiper Belt objects beyond a distance of
50 AU from the Sun (Jewitt et al. 1998; Trujillo & Brown
2001; Petit et al. 2006). However, the origin of this edge
is unclear. Adams et al. (2004) note that the observed
distance is far interior to the radius at which truncation
by photoevaporation would be expected to occur, while
Youdin & Shu (2002) find that the presence of such an
edge in planetesimal density could be explained by drift-
induced enhancement. A compelling possibility is that
the Sun formed in a cluster environment, and the early
solar disk was truncated by a close encounter with a pass-
ing star (see Reipurth 2005, and references therein). A
more complete understanding of the outer regions of pro-
toplanetary disks may provide insight into the processes
that shape the outer solar system.
Pi´etu et al. (2005) present multiwavelength millime-
ter continuum and CO isotopologue observations of the
disk around the Herbig Ae star AB Aurigae and found
from fitting models of disk structure described by trun-
cated power laws that the outer radius of the dust derived
from continuum emission (350±30 AU) was much smaller
than that of the gas derived from 12CO J=2-1 emission
(1050±10 AU). They suggest that a change in dust grain
properties resulting in a drop in opacity could be respon-
sible for the difference, and note the possible association
with a ring feature in the disk at 200 AU. A similar result
was obtained by Isella et al. (2007) from observations of
the disk around the Herbig Ae star HD 163296: they
found a significant discrepancy between the outer radius
derived for the dust continuum emission (200 ± 15 AU)
and that derived from CO emission (540± 40 AU). These
data appeared to require a sharp drop in surface density,
opacity, or dust-to-gas ratio beyond 200 AU; however, as
they discuss, there is no obvious physical basis for such
a discontinuity. As Isella et al. (2007) demonstrate, the
2
discrepancy in outer radii derived from the dust and gas
is not simply an issue of sensitivity; the observations were
sufficiently sensitive to detect emission from the power-
law dust disk if it did extend to the radius indicated by
the CO emission. The underlying issue is that the trun-
cated power law model does not simultaneously repro-
duce the extent of both the continuum and CO emission
for these disks.
Using data from the Submillimeter Array3 we show
that the same apparent discrepancy in gas and dust outer
radius applies to the circumstellar disks around several
more young stars. In an attempt to understand the origin
of this discrepancy, we investigate an alternative surface
density profile based on work by Hartmann et al. (1998),
which is similar to a power law profile in the inner disk
but includes a tapered outer edge. We show that this
model, which has a physical basis in similarity solutions
of disk evolution with time, is capable of simultaneously
reproducing both continuum and CO emission from these
disks. The primary difference between this model and the
truncated power-law disk is that instead of a sharp outer
edge the surface density falls off gradually, with sufficient
column density at large radii that CO emission extends
beyond the point at which dust continuum emission be-
comes negligible.
2. MILLIMETER/SUBMILLIMETER DUST CONTINUUM
AND CO J=3-2 DATA
The analysis was conducted on extant SMA data of
the disks around of HD 163296, TW Hydrae, GM Au-
rigae, and MWC 480. The dates, frequencies, antenna
configurations, number of antennas, and original publi-
cations associated with the data sets are listed in Table
1. The four disk systems chosen for this analysis are all
nearby, bright, isolated, and have been well studied at a
wide range of wavelengths. The velocity fields of these
disks all appear to be well described by Keplerian rota-
tion (Isella et al. 2007; Qi et al. 2004; Dutrey et al. 1998;
Pi´etu et al. 2007). The relevant properties of these sys-
tems (spectral type, distance, stellar mass, age, and disk
inclination and position angle) are listed in Table 2.
3. DISK MODELS
Using the SMA data available for the four disk systems,
we compared two classes of disk models: the first model
is described by power laws in surface density and temper-
ature and is truncated at an outer radius Rout (details
in §3.1), and the second model is described by a power
law in temperature and a surface density profile similar
to a power law in the inner disk but tapered with an ex-
ponential edge in the outer disk (details in §3.2). This
latter model is not intended to be a definitive description
of these disks, but rather illustrative of the broader cat-
egory of models without a sharp outer edge. The model
fitting process involved deriving a minimum χ2 solution
for those parameters of each class of model that best
fit the continuum emission, and then using standard as-
sumptions to predict CO emission (described in §3.4).
The CO emission was not used to determine the model
3 The Submillimeter Array is a joint project between the Smith-
sonian Astrophysical Observatory and the Academia Sinica Insti-
tute of Astronomy and Astrophysics and is funded by the Smith-
sonian Institution and the Academia Sinica.
fits, due to the computational intensity of solving the ex-
citation and radiative transfer for the molecular line for
a large grid of models.
3.1. Truncated Power Law
For the truncated power law models, we used the pre-
scription of Dutrey et al. (1994). In this framework, the
disk structure is described by power laws in tempera-
ture and surface density, with the scale height specified
through the assumption that the disk is in hydrostatic
equilibrium:
100AU(cid:19)−q
T (R) = T100(cid:18) R
100AU(cid:19)−p
Σ(R) = Σ100(cid:18) R
H(R) =s 2R3kBTk(R)
GM⋆m0
(1)
(2)
(3)
where the subscript '100' refers to the value at 100 AU,
kB is Boltzmann's constant, G is the gravitational con-
stant, M⋆ is the stellar mass, and m0 is the mass per
particle (we assume 2.37 times the mass of the hydrogen
atom). Combining these expressions and the assumption
of hydrostatic equilibrium, the volume density n(R, z) is
given by:
n(R, z) =
Σ(R)
√πH(R)
exp −(z/H(R))2
(4)
where z is the vertical height above the midplane. As
noted by Dutrey et al. (2007), this definition implies a
scale height of H(r) = √2cs/Ω, where cs is the sound
speed and Ω the angular velocity, while other groups use
H(r) = cs/Ω; this difference should be taken into ac-
count when comparing our results with other disk struc-
ture models. During the modeling process, we recast the
surface density normalization in terms of the midplane
density at 100 AU, so that the parameter Σ100 is replaced
by n100. This power-law model of disk structure has five
free parameters: T100, q, n100, p, and Rout.
3.2. Similarity Solution from Accretion Disk Evolution
While versatile and ubiquitous, the truncated power
law models of disk structure have one obviously unphys-
ical feature: a sharp outer edge. In the absence of dy-
namical effects (e.g. from a binary companion) or large
pressure gradients to confine the material, disk structure
at the outer edge is expected to taper off gradually. A de-
scription of the structure of an isolated, steadily accret-
ing disk as it evolves with time is provided by Hartmann
et al. (1998), who expand on the work of Lynden-Bell &
Pringle (1974) to show that if the viscosity in a disk can
be written as a time-independent power law of the form
ν ∝ Rγ, then the similarity solution for the disk surface
density is given by
Σ(r) =
C
rγ T −(5/2−γ)/(2−γ) exp(cid:20)−
r2−γ
T (cid:21)
(5)
where C is a constant, r is the disk radius in units of
the radial scale factor R1 such that r = R/R1, and T
3
Sources of SMA 230/345 GHz continuum and CO J=3-2 data.
TABLE 1
Freq./
Transition Object
230 GHz
HD 163296
TW Hydrae
GM Aurigae
MWC 480
345 GHz/ HD 163296
CO J=3-2 TW Hydrae
GM Aurigae
MWC 480
Dates
Array
Config.
No. of
Antennas Reference
23/24 Aug 2003 Compact N
Extended
Compact
Extended
18/20 Nov 2003 Compact N
10 Apr 2005
27 Feb 2005
10 Dec 2006
23 Aug 2005
28 Dec 2006
26 Nov 2005
21 Oct 2005
Compact
Compact N
Compact
Compact
7
8
8
8
8
8
8
7
8
1
2
2
3
1
4
5
6
1
References. -- (1) SMA archive; (2) Qi et al. (2006); (3) Qi et al. (in prep); (4) Isella et al. (2007) ; (5) Qi et al. (2007, submitted);
(6) Andrews & Williams (2007).
is the nondimensional time T = t/ts + 1 where ts is the
viscous scaling time (eq. 20 in Hartmann et al. 1998).
For simplicity, when applying these models to our data
we used physical units and absorbed several of the pa-
rameters into two constants so that the surface density
is of the form
Σ(R) =
c1
Rγ exp"−(cid:18) R
c2(cid:19)2−γ# ,
(6)
where R is the disk radius in AU and c1, c2, and γ are
constants that we allowed to vary during the fitting pro-
cess.
The temperature profile for the similarity solution disk
model is identical to that of the truncated power-law
disk, except that its spatial extent is infinite. We do not
allow it to drop below 10 K, but this low temperature
limit is not problematic for any of the disks considered
here. This model therefore includes five free parameters:
T100, q, c1, γ, and c2. The constant c1 describes the
normalization of the surface density, similar to n100 in
the power-law model, while the constant c2 is analogous
to the outer radius, since it describes the radial scale
length over which the exponential taper acts to cause
the surface density to drop towards zero.
3.3. Model Comparison
The surface density description for the similarity solu-
tion is similar to the truncated power law except at the
outer edge of the disk. In the inner regions of the disk for
which R ≪ c2, we may expand the exponential so that
exp(cid:2)−(R/c2)2−γ(cid:3) → 1 − ( R
)2−γ + · · ·, and the surface
density becomes
c2(cid:19)2−γ
Rγ (1 −(cid:18) R
R2(1−γ) (7)
Σ(R) =
c1
c2
) =
c1
Rγ −
c1
c2−γ
2
In the α-viscosity context (Shakura & Syunyaev 1973),
for a vertically isothermal disk with the typical temper-
ature index q = 0.5, we would expect that γ = 1. This
implies that for standard assumptions, the inner disk sur-
face density will be described by a power law in R with
index γ, modified by a constant ( c1
) due to the influence
c2
of the exponential. If γ deviates from 1, an additional
shallow dependence on R would be expected.
It is illuminating to consider the behavior of these mod-
els in the Fourier domain, the natural space for interfer-
ometer observations. To do so we define the coordinate
Ruv, the distance from the phase center of the disk in
the (u, v) plane, as it would be observed if the disk were
viewed directly face-on. To perform the deprojection
from the inclined and rotated sky coordinates, we cal-
culate the position of each point in the (u, v) plane as a
projected distance from the major and minor axes of the
disk, respectively: da = R sin φ and db = R cos φ cos i,
where i is the disk inclination, R = (u2 + v2)1/2, φ
is the polar angle from the major axis of the disk,
φ = arctan(v/u − P A), and P A is the position angle
measured east of north; then Ruv = (d2
b )1/2 (Lay
et al. 1997; Hughes et al. 2007).
In the Fourier domain, the truncated disk models show
"ringing" and the visibilities will drop rapidly to zero in
the vicinity of Ruv = 1/Rout. Under the simplifying
assumption of γ = 1, the Fourier transform of the sim-
ilarity solution becomes a convolution of two functions:
(1) 1/R, which is just the Fourier transform of the 1/R
dependence of a power law extending from zero to infin-
ity, and (2) c2/(1 + R2c2)3/2, where c is a scaling con-
stant for the term that describes the exponential taper.
Since these two functions both decrease monotonically
and are always positive, the visibilities drop smoothly to
zero without any ringing.
a + d2
3.4. Model Fitting
For both disk models, we fit for the five parameters de-
scribing the disk temperature and surface density struc-
ture using the continuum data for each disk with the
widest range of available baseline lengths. The position
angle and inclination were fixed and adopted from pre-
vious studies (see Table 2). For opacity, we assume the
standard millimeter value adopted by Beckwith & Sar-
gent (1991) (κν = κ0(ν/ν0)β, where κ0 = 10.0 cm2/gdust,
ν0=1 THz, and β = 1), although we allow β to vary in or-
der to obtain the proper normalization when extrapolat-
ing from one frequency to another. Due to the ∼ 100 AU
spatial resolution, these data are not sensitive to the in-
ner radius of the disk. For both sets of models, therefore,
we simply fix the inner radius at a value of 4 AU for TW
Hya (Calvet et al. 2002; Hughes et al. 2007), 24 AU for
GM Aur (Calvet et al. 2005), and 3 AU for the other
two systems, for which reliable inner radius information
is not available; this is sufficiently small that changes in
the inner radius do not affect the derived model param-
eters. To compare the models to the data, we use the
Monte-Carlo radiative transfer code RATRAN (Hoger-
heijde & van der Tak 2000) to calculate sky-projected
4
images of the dust continuum and CO emission, with
frequency and bandwidth appropriate for the observa-
tions, and assuming Keplerian rotation. We then use
the MIRIAD task uvmodel to simulate the SMA obser-
vations, with the same antenna positions and visibility
weights.
For each set of parameters, we directly compare the
model visibilities to the continuum data and calculate
a χ2 value, using the minimum χ2 value to determine
the best-fit parameters. The resulting best-fit models
are shown along with continuum data for both frequen-
cies in the left panels of Figure 1. The abscissa gives
the deprojected radial distance in the (u, v) plane, and
the ordinate shows the real and imaginary components
of the visibility. For a disk with circular symmetry, the
imaginary components should average to zero. The 230
GHz continuum data are depicted as open circles, while
the 345 GHz data are filled circles. The best-fit power-
law model is shown in blue and the similarity solution in
orange. Dotted and dashed lines distinguish between the
230 and 345 GHz model predictions; the fit was deter-
mined at that frequency with the largest baseline cover-
age and extrapolated to the other frequency, by varying
β. The uncertainties quoted for β reflect an assumed
10% calibration uncertainty. Note that varying β has no
effect on the modeled CO emission.
We measure values of β that are consistent with 1,
which is in agreement with the typical values measured
for disks in the Taurus-Auriga association (e.g. Dutrey
et al. 1996; Rodmann et al. 2006). These shallow mil-
limeter spectral slopes indicate that some grain growth
has occurred from ISM grain sizes, which typically ex-
hibit a steeper spectral slope (β ∼ 2). In particular, the
value of 1.2 measured for GM Aur matches well the value
of 1.2 reported by Andrews & Williams (2007), and the
value of 0.7 measured for TW Hydrae matches well the
value of 0.7 reported by Calvet et al. (2002) and Natta
et al. (2004).
After fitting the continuum, we then assumed a
gas/dust mass ratio of 100 and a standard interstel-
lar CO/H2 mass ratio of 10−4 to predict the expected
strength and spatial extent of CO emission from the
disks, based on the best-fit continuum model. We as-
sume throughout that the gas and dust are well-mixed,
and that CO traces molecular hydrogen. We do not take
into account the complexities of disk chemistry, such as
the depletion of CO molecules in the cold, dense mid-
plane (Aikawa 2007; Semenov et al. 2006). However, de-
viations from these simple assumptions should have no
appreciable effect on our conclusions concerning the ra-
dial extent of CO emission.
Since we neglect the vertical temperature gradient in
the disk, we might expect to underpredict the strength
of optically thick CO line emission, which likely origi-
nates in the upper layers of the disk that are subject to
heating by stellar irradiation. The continuum emission,
by contrast, is likely weighted toward the cold midplane
of the disk. For this reason, after obtaining an initial fit
from the continuum, we allowed the temperature scale
(T100) to vary to best reproduce the flux levels of the ob-
served CO emission, and then iteratively fit for the other
structural parameters (q, n100, c2/Rout, and γ/p).
Deriving the temperature from the CO emission in this
way may underestimate the midplane density in some
cases, due to the degeneracy between T100 and n100:
the temperature derived from CO emission is typically
greater than or equal to that of the shielded midplane,
depending in detail on the dust opacity and molecular
dissociation due to ultraviolet radiation in the upper
disk layers (for a discussion of the processes involved,
see Jonkheid et al. 2007; Isella et al. 2007). For the disks
considered, the temperature derived for the dust contin-
uum emission was within ∼ 40% of that derived to match
the CO line strength.
4. RESULTS AND DISCUSSION
The parameters for the best-fit model solutions to the
continuum data for each source and for each of the two
model types are listed in Table 3. This table lists only
the set of parameters with the minimum χ2 value; for-
mal errors are not quoted as these are not intended to
be definitive structural models but simply illustrative of
the differences between the model classes in their treat-
ment of the outer edge. The midplane surface density
profiles for these models are plotted in Figure 2. The
solid lines depict the profile for the power-law solution,
while the dashed lines are for the similarity solution. The
parameters of the two model solutions are very similar,
particularly for HD 163296 and MWC 480. For all four
disks, the two model solutions are particularly similar
just within the outer edge of the disk, around the range
of radii well-matched to the resolution of the data (∼ 200
AU for HD 163296, ∼ 90 AU for TW Hydrae, ∼ 200 AU
for GM Aurigae, and ∼ 300 AU for MWC 480). The
outer radius for the power-law solution typically falls at
roughly twice the scale length (c2) of the similarity solu-
tion. The analogous parameters γ and p, which describe
how quickly the midplane density drops with radius, are
also very similar between the two models.
CO J=3-2 emission predicted from these best-fit mod-
els is shown in the right panel of Figure 1. The similar-
ity solution is shown in the blue-contoured central plot,
and the power-law model in the orange-contoured plot
on the right. Recessional velocity is plotted on the ab-
scissa while the position offset along a slice through the
disk major axis is plotted on the ordinate. The horizon-
tal dashed line in each figure represents the extent of the
outer radius (Rout) derived for that source in the context
of the truncated power-law model. For all four sources,
the extent of molecular gas emission from the similarity
solution is much more closely matched to the data than
that of the power-law model, even though both reproduce
the continuum dust emission equally well.
From Figure 1,
it is clear by eye that for all four
sources, the extent of the CO emission is severely under-
predicted by the power law model but matches well the
predicted emission from the similarity solution model.
A calculation of the χ2 value comparing the predicted
CO emission for the two models to the observed emis-
sion shows that the similarity solution matches the data
better than the truncated power-law model for all of the
disks in our study. The difference is at the 2σ level for
MWC 480, for which there is only short-baseline data
with relatively low signal-to-noise, and at the 4σ level for
GM Aur; for TW Hydrae and HD 163296, the χ2 analy-
sis shows that, formally, the similarity solution provides
a better fit to the CO emission than the power-law model
5
TABLE 2
Stellar and disk properties
Spectral Dist.
Stellar
System
HD 163296
TW Hydrae
GM Aurigae
MWC 480
Type
A1V
K8V
K5V
A3
(pc) Mass (M⊙)
122a
51d,e
140
140m
2.3a
0.6
0.8i
1.8n
Age
(Myr)
5b
5-20f,g
2-10j,k
7-8n,o
Disk
PA (◦)
145c
-45h
51i
143p
Disk
i (◦)
46c
7h
56i
37p
References. -- (a) van den Ancker et al. (1998); (b) Natta et al. (2004); (c) Isella et al. (2007); (d) Mamajek (2005); (e) Hoff et al.
(1998); (f ) Kastner et al. (1997); (g) Webb et al. (1999); (h) Qi et al. (2004); (i) Dutrey et al. (1998); (j) Beckwith et al. (1990); (k) Simon
& Prato (1995); (m) The et al. (1994); (n) Pi´etu et al. (2007); (o) Simon et al. (2000); (p) Hamidouche et al. (2006).
Fig. 1. -- Comparison between the data and the two types of models (similarity solution and power law) for the four disks in our sample:
(a) HD 163296, (b) TW Hydrae, (c) GM Aurigae, and (d) MWC 480. For each source, the left panel shows the real and imaginary
visibilities as a function of deprojected (u, v) distance from the phase center. Symbols are SMA data; open circles are 230 GHz and filled
circles are 345 GHz continuum. The lines represent the best fit to the 345 GHz continuum for the power law (orange) and similarity (blue)
models. Dashed lines show the model at 345 GHz while solid lines are 230 GHz. The right panel shows position-velocity diagrams of the
J=3-2 rotational transition of CO along the major axis of the disk. The left plot (black contours) shows the SMA data. The middle plot
(blue contours) displays the emission predicted by the similarity solution parameters that provide the best fit to the continuum emission,
and the right plot (orange contours) displays the emission predicted for the best-fit power-law model. The horizontal dashed line across
the right panel represents the extent of the outer radius (Rout) derived for each source through fitting of the continuum emission in the
context of the power-law model. The contour levels, beam, and velocity resolution for each source are as follows: (a) [2,4,6,8,10,12]×1.1
Jy/beam, 3.0×2.1 arcsec at a position angle of 14.3◦, and 0.35 km/s; (b) [2,4,6,8]×2.0 Jy/beam, 4.0×1.8 arcsec at a position angle of 3.2◦,
and 0.18 km/s; (c) [2,4,8,12,16]×0.5 Jy/beam, 2.3×2.1 arcsec at a position angle of 12.9◦, and 0.35 km/s; (d) [2,4,6,8,10]×0.5 Jy/beam,
2.5×2.3 arcsec at a position angle of 45.3◦, and 0.35 km/s.
6
Parameters for best-fit continuum models
TABLE 3
Source
HD 163296
TW Hydrae
GM Aurigae
MWC 480
χ2
Model
Similarity
2.29
Power Law 2.26
Similarity
2.42
Power Law 2.41
Similarity
2.19
Power Law 2.17
Similarity
1.86
Power Law 1.86
T100 (K)
65
60
40
30
50
40
50
45
q
0.4
0.5
0.2
0.5
0.5
0.4
0.8
0.7
n10
a (cm−3) Rout (AU)
c2 (AU)
5.3 × 1011
6.7 × 1011
2.3 × 1011
7.1 × 1010
1.1 × 1011
5.0 × 1011
1.0 × 1011
1.3 × 1011
125
250
30
60
140
275
200
275
γ
p
0.9
1.0
0.7
1.0
0.9
1.3
1.1
1.3
β
0.4+0.5
−0.3
0.5+0.5
−0.3
0.7+0.5
−0.1
0.7+0.5
−0.1
1.2+0.5
−0.1
1.3+0.5
−0.1
0.7+0.5
−0.4
0.7+0.5
−0.4
aMidplane density at 10 AU. We use the value at 10 AU rather
than 100 AU to compare better the power law and similarity models
in the region where their behavior is similar.
7
it to do so, it illustrates that the outer radius discrep-
ancy is peculiar to the truncated power-law model; other
disk structure models with a tapered outer edge may be
able to reproduce the gas and dust emission as well as,
or better than, the similarity solution adopted here.
Analysis of the CO excitation in the similarity solution
model shows that the extent of the CO J=3-2 emission
in these disks coincides roughly with the radius at which
the line excitation becomes subthermal, determined pri-
marily by where the mid-plane density drops below the
critical density (∼ 4.4 × 104 cm−3 at 20 K, though effec-
tively lowered when photon trapping plays a role). In the
similarity solution model, the surface density distribu-
tion steepens dramatically at large radii, but without the
sharp truncation of the power-law model. This suggests
that caution should be exercised not only when compar-
ing outer radius measurements based on dust continuum
and molecular gas emission, but also when comparing
measurements based on emission from different transi-
tions of CO or from isotopologues of the CO molecule
that have differing abundances and optical depths. Pi´etu
et al. (2007) fit truncated power law models to the disks
around DM Tau, LkCa 15, and MWC 480 in several dif-
ferent isotopologues and rotational transitions of CO.
For the two cases in which multiple transitions of the
13CO molecule were observed, the derived outer radius
is marginally smaller for the J=2-1 transition than the
J=1-0 transition. This result is consistent with the ex-
pected trend that lower-J transitions will exhibit larger
outer radii due to their lower critical density: a lower crit-
ical density will be reached at a greater distance as the
surface density tapers off near the outer edge of the disk.
In all cases the Pi´etu et al. (2007) analysis also yielded
a smaller outer radius in 13CO than in 12CO, as well as
a flatter surface density power law index for 13CO than
for 12CO. These differences may be related to selective
photodissociation, or other chemical processes. However,
the trends of smaller outer radius and shallower surface
density index in 13CO are also consistent with surface
density falling off rapidly at large radii, as expected for a
disk with a tapered outer edge. In the similarity solution
model, the less abundant 13CO isotopologue will become
undetectable at smaller radii than 12CO, which is more
sensitive to the exponential drop in surface density in the
outer disk.
It is noteworthy that studies of six largest "proplyds"
with the most distinct silhouettes in the Orion Nebula
Cluster reveal radial profiles in extinction that are well-
described by an exponential taper at the outer edge (Mc-
Caughrean & O'Dell 1996). These isolated disks may be
analogous to the systems considered here with a tapered
outer edge.
Models with tapered outer edges also aid in addressing
discrepancies between the size of the dust disk observed
in the millimeter and the extent of scattered light ob-
served in the optical and near-infrared. For example,
coronographic observations of TW Hydrae detect scat-
tered light to a distance of ∼ 200 AU from the star
(Krist et al. 2000; Trilling et al. 2001; Weinberger et al.
2002), while the truncated power-law model places the
outer edge of the dust disk closer to 60 AU. Similarly,
observations of HD 163296 by Grady et al. (2000) de-
tect scattered light out to ∼ 400 AU from the star, much
Fig. 2. -- Midplane density structure of the models that provide
the best fit to the continuum data. Solid lines show truncated
power-law models while dashed lines show similarity solution mod-
els.
at the > 10σ level.
The tapered edge of the similarity solution density dis-
tribution evidently permits a large enough column den-
sity to produce detectable CO 3-2 line emission, even
though it has dropped off enough that the continuum
emission is negligible. The power-law model, by con-
trast, is strictly limited in the extent of its CO emission
by the sharp outer radius. In particular, for the case of
HD 163296, the CO emission predicted by the power law
model (orange contours in the right panel of Figure 1a)
falls to 4.4 Jy/beam at a distance of 1.8 arcsec (220 AU)
from the source center, while the similarity solution (blue
contours) maintains this brightness out to a distance of
4.7 arcsec (570 AU). This latter size is well matched to
the data (black contours) which extends at this bright-
ness to a distance of 5.0 arcsec (600 AU). These dis-
tances likely overestimate the true physical extent of the
disk due to convolution with the 2.1 × 3.0 arcsec beam,
though they are very comparable to the values observed
by Isella et al. (2007). While the similarity solution does
not provide a perfect fit to the data, nor do we intend
8
larger than the 250 AU radius of the dust disk implied
by the truncated power-law model. While the exponen-
tial taper causes the density of the similarity solution to
drop rapidly with radius, these models retain a substan-
tial vertical column density for several exponential scale
lengths. It is therefore plausible that scattered light can
remain visible at this distance, in contrast to the case of
the smaller truncated power-law disk.
Although we intend for the similarity solution applied
here to be an illustrative rather than definitive descrip-
tion of the disk structure, it is important to note that
the particular form applied here has potential implica-
tions for the study of the evolutionary status of these
disks. The form of the similarity solution developed by
Lynden-Bell & Pringle (1974) and Hartmann et al. (1998)
relates the observed structure to the disk age, viscosity,
and initial radius. Although all three of these variables
are poorly constrained by current observations, a large
and homogeneous sample of objects studied in this way
might reveal evolutionary trends in the disk structure.
5. SUMMARY AND CONCLUSIONS
With the advent of high signal-to-noise interferometer
observations that resolve the outer regions of nearby pro-
toplanetary disks, an apparent discrepancy has emerged
between the extent of the dust continuum and molecular
gas emission (Pi´etu et al. 2005; Isella et al. 2007). Us-
ing multi-frequency interferometric data from the Sub-
millimeter Array, we have investigated this disparity for
four disk systems (HD 163296, TW Hydrae, GM Auri-
gae, and MWC 480) in the context of two distinct classes
of disk structure models: (1) a truncated power law, and
(2) a similarity solution for the time evolution of an accre-
tion disk. The primary difference between these models
is in their treatment of the disk outer edge: the abruptly
truncated outer edge of the power-law disk causes the
visibilities to drop rapidly to zero, leading to an inferred
outer radius that is small in comparison with the ob-
served molecular gas emission. The similarity solution,
by contrast, tapers off smoothly, creating a broader vis-
ibility function and allowing molecular gas emission to
persist at radii well beyond the region in the disk where
continuum falls below the detection threshold. The outer
radius discrepancy appears to exist only in the context
of the power-law models.
In light of this result, it appears that an abrupt change
in dust properties for these disks is unlikely, as there is
no physical mechanism to explain such a discontinuity.
This may imply that a sharp change in dust properties in
the early solar nebula is similarly an unlikely explanation
for the Kuiper belt edge observed by Jewitt et al. (1998),
and that a dynamical mechanism such as truncation by
a close encounter with a cluster member (Reipurth 2005,
and references therein) may provide a more plausible ori-
gin. In this case, we would expect to observe disks with
sharp outer edges only in clustered environments, and
a model with a tapered edge would be a more realistic
prescription for investigating the structure of a typical
isolated disk. The tapered disk models provide a natural
explanation for the disparate outer radii observed using
different probes of the disk extent, including comparison
of continuum and molecular gas observations (Pi´etu et al.
2005; Isella et al. 2007), and also comparison of different
isotopologues and rotational transitions of a particular
molecule (Pi´etu et al. 2007). When predicting CO emis-
sion, this simple model does neglect potential variance
in the CO abundance due to depletion in the midplane
and photodissociation at the disk surface; however, the
results presented are intended simply to illustrate the
global differences between gas and dust emission from
the two model classes, independent of detailed CO chem-
istry.
While we cannot rule out disparate gas and dust radii
in these disks, we show that an alternative disk struc-
ture model, grounded in the physics of accretion, resolves
the apparent size discrepancy without the need to invoke
dramatic changes in dust opacity, dust density, or dust-
to-gas ratio in the outer disk.
The authors would like to thank Sean Andrews
for thoughtful comments that helped to improve the
manuscript. Partial support for this work was pro-
vided by NASA Origins of Solar Systems Program Grant
NAG5-11777. A. M. H. acknowledges support from a
National Science Foundation Graduate Research Fellow-
ship.
REFERENCES
Adams, F. C., Hollenbach, D., Laughlin, G., & Gorti, U. 2004,
Hamidouche, M., Looney, L. W., & Mundy, L. G. 2006, ApJ, 651,
ApJ, 611, 360
Adams, F. C., Lada, C. J., & Shu, F. H. 1987, ApJ, 312, 788
Adams, F. C. & Shu, F. H. 1986, ApJ, 308, 836
Aikawa, Y. 2007, ApJ, 656, L93
Andrews, S. M. & Williams, J. P. 2007, ApJ, 659, 705
Beckwith, S. V. W. & Sargent, A. I. 1991, ApJ, 381, 250
Beckwith, S. V. W., Sargent, A. I., Chini, R. S., & Guesten, R.
1990, AJ, 99, 924
321
Hartmann, L., Calvet, N., Gullbring, E., & D'Alessio, P. 1998,
ApJ, 495, 385
Hoff, W., Henning, T., & Pfau, W. 1998, A&A, 336, 242
Hogerheijde, M. R. & van der Tak, F. F. S. 2000, A&A, 362, 697
Hughes, A. M., Wilner, D. J., Calvet, N., D'Alessio, P., Claussen,
M. J., & Hogerheijde, M. R. 2007, ApJ, 664, 536
Isella, A., Testi, L., Natta, A., Neri, R., Wilner, D., & Qi, C.
Calvet, N., D'Alessio, P., Hartmann, L., Wilner, D., Walsh, A., &
2007, A&A, 469, 213
Sitko, M. 2002, ApJ, 568, 1008
Calvet, N., et al. 2005, ApJ, 630, L185
Dutrey, A., Guilloteau, S., Duvert, G., Prato, L., Simon, M.,
Schuster, K., & Menard, F. 1996, A&A, 309, 493
Dutrey, A., Guilloteau, S., & Ho, P. 2007, in Protostars and
Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil, 495 -- 506
Dutrey, A., Guilloteau, S., Prato, L., Simon, M., Duvert, G.,
Schuster, K., & Menard, F. 1998, A&A, 338, L63
Dutrey, A., Guilloteau, S., & Simon, M. 1994, A&A, 286, 149
Grady, C. A., et al. 2000, ApJ, 544, 895
Jewitt, D., Luu, J., & Trujillo, C. 1998, AJ, 115, 2125
Jonkheid, B., Dullemond, C. P., Hogerheijde, M. R., & van
Dishoeck, E. F. 2007, A&A, 463, 203
Kastner, J. H., Zuckerman, B., Weintraub, D. A., & Forveille, T.
1997, Science, 277, 67
Krist, J. E., Stapelfeldt, K. R., M´enard, F., Padgett, D. L., &
Burrows, C. J. 2000, ApJ, 538, 793
Lay, O. P., Carlstrom, J. E., & Hills, R. E. 1997, ApJ, 489, 917
Lay, O. P., Carlstrom, J. E., Hills, R. E., & Phillips, T. G. 1994,
ApJ, 434, L75
Lynden-Bell, D. & Pringle, J. E. 1974, MNRAS, 168, 603
Mamajek, E. E. 2005, ApJ, 634, 1385
McCaughrean, M. J. & O'Dell, C. R. 1996, AJ, 111, 1977
Mundy, L. G., McMullin, J. P., Grossman, A. W., & Sandell, G.
1993, Icarus, 106, 11
Natta, A., Testi, L., Neri, R., Shepherd, D. S., & Wilner, D. J.
2004, A&A, 416, 179
Petit, J.-M., Holman, M. J., Gladman, B. J., Kavelaars, J. J.,
Scholl, H., & Loredo, T. J. 2006, MNRAS, 365, 429
Pi´etu, V., Dutrey, A., & Guilloteau, S. 2007, A&A, 467, 163
Pi´etu, V., Guilloteau, S., & Dutrey, A. 2005, A&A, 443, 945
Qi, C., et al. 2004, ApJ, 616, L11
Qi, C., Wilner, D. J., Calvet, N., Bourke, T. L., Blake, G. A.,
Hogerheijde, M. R., Ho, P. T. P., & Bergin, E. 2006, ApJ, 636,
L157
Reipurth, B. 2005, in Astronomical Society of the Pacific
Conference Series, Vol. 341, Chondrites and the Protoplanetary
Disk, ed. A. N. Krot, E. R. D. Scott, & B. Reipurth, 54 -- +
Rodmann, J., Henning, T., Chandler, C. J., Mundy, L. G., &
Wilner, D. J. 2006, A&A, 446, 211
9
Semenov, D., Wiebe, D., & Henning, T. 2006, ApJ, 647, L57
Shakura, N. I. & Syunyaev, R. A. 1973, A&A, 24, 337
Simon, M., Dutrey, A., & Guilloteau, S. 2000, ApJ, 545, 1034
Simon, M. & Prato, L. 1995, ApJ, 450, 824
The, P. S., de Winter, D., & Perez, M. R. 1994, A&AS, 104, 315
Trilling, D. E., Koerner, D. W., Barnes, J. W., Ftaclas, C., &
Brown, R. H. 2001, ApJ, 552, L151
Trujillo, C. A. & Brown, M. E. 2001, ApJ, 554, L95
van den Ancker, M. E., de Winter, D., & Tjin A Djie, H. R. E.
1998, A&A, 330, 145
Webb, R. A., Zuckerman, B., Platais, I., Patience, J., White,
R. J., Schwartz, M. J., & McCarthy, C. 1999, ApJ, 512, L63
Weidenschilling, S. J. 1977, Ap&SS, 51, 153
Weinberger, A. J., et al. 2002, ApJ, 566, 409
Youdin, A. N. & Shu, F. H. 2002, ApJ, 580, 494
|
astro-ph/0204180 | 1 | 0204 | 2002-04-10T18:47:29 | Photoevaporation of Clumps in Photodissociation Regions | [
"astro-ph"
] | We present the results of an investigation of the effects of Far Ultraviolet (FUV) radiation from hot early type OB stars on clumps in star-forming molecular clouds. Clumps in Photodissociation regions (PDRs) undergo external heating which, if rapid, creates strong photoevaporative mass flows off the clump surfaces, and drives shocks into the clumps, compressing them to high densities. The clumps lose mass on relatively short timescales. The evolution of an individual clump is found to be sensitive to its initial colunm density, the temperature of the heated surface and the ratio of the ``turn-on time'' $t_{FUV}$ of the heating flux on a clump to its initial sound crossing-time $t_{c}$.
In this paper, we use spherical 1-D numerical hydrodynamic models as well as approximate analytical models to study the evolution of turbulence-generated and pressure-confined clumps in PDRs. Turbulent clumps evolve so that their column densities are equal to a critical value determined by the local FUV field, and typically have short photoevaporation timescales, $\sim 10^{4-5}$ years for a 1 M$_{\odot}$ clump in a typical star-forming region. Clumps that are confined by an interclump medium may either get completely photoevaporated, or may preserve a shielded core with a warm, dissociated, protective shell that absorbs the incident FUV flux. We compare our results with observations of some well-studied PDRs: the Orion Bar, M17SW, NGC 2023 and the Rosette Nebula. The data are consistent with both interpretations of clump origin, with a slight indication for favouring the turbulent model for clumps over pressure-confined clumps. | astro-ph | astro-ph |
Photoevaporation of Clumps in Photodissociation Regions
U.Gorti and D.Hollenbach
NASA Ames Research Center, Moffett Field, CA
ABSTRACT
We present the results of an investigation of the effects of Far Ultraviolet
(FUV) radiation (6.0eV < hν < 13.6eV ) from hot early type OB stars on clumps
in star-forming molecular clouds. Clumps in FUV-illuminated regions (or pho-
todissociation regions or PDRs) undergo external heating and photodissociation
as they are exposed to the FUV field, resulting in a loss of cold, molecular clump
mass as it is converted to warm atomic gas. The heating, if rapid, creates strong
photoevaporative mass flows off the clump surfaces, and drives shocks into the
clumps, compressing them to high densities. The clumps lose mass on relatively
short timescales. The evolution of an individual clump is found to be sensitive
to three dimensionless parameters: ηc0, the ratio of the initial column density of
the clump to the column N0 ∼ 1021 cm−2 of a warm FUV-heated surface region;
ν, the ratio of the sound speed in the heated surface to that in the cold clump
material; and tF U V /tc, the ratio of the "turn-on time" tF U V of the heating flux
on a clump to its initial sound crossing-time tc. The evolution also depends on
whether a confining interclump medium exists, or whether the interclump region
has negligible pressure, as is the case for turbulence-generated clumps. In this
paper, we use spherical 1-D numerical hydrodynamic models as well as approx-
imate analytical models to study the dependence of clump photoevaporation on
the physical parameters of the clump, and to derive the dynamical evolution,
mass loss rates and photoevaporative timescales of a clump for a variety of as-
trophysical situations. Turbulent clumps evolve so that their column densities
are equal to a critical value determined by the local FUV field, and typically
have short photoevaporation timescales, ∼ 104−5 years for a 1 M⊙ clump in a
typical star-forming region (ηc0 = 10, ν = 10). Clumps with insufficient magnetic
pressure support, and in strong FUV fields may be driven to collapse by the com-
pressional effect of converging shock waves. We also estimate the rocket effect
on photoevaporating clumps and find that it is significant only for the smallest
clumps, with sizes much less than the extent of the PDR itself. Clumps that are
confined by an interclump medium may either get completely photoevaporated,
or may preserve a shielded core with a warm, dissociated, protective shell that
-- 2 --
absorbs the incident FUV flux. We compare our results with observations of some
well-studied PDRs: the Orion Bar, M17SW, NGC 2023 and the Rosette Nebula.
The data are consistent with both interpretations of clump origin, turbulence and
pressure confinement, with a slight indication for favouring the turbulent model
for clumps over pressure-confined clumps.
Subject headings: ISM:clouds -- ISM:dynamics -- stars:early-type -- stars:formation
-- ultraviolet:stars
1.
Introduction
Young massive OB stars significantly influence the structure, dynamics, chemistry and
thermal balance of their associated molecular clouds through the impact of their ultraviolet
photons. Their extreme ultraviolet photons (EUV, hν > 13.6 eV) ionize the gas immediately
surrounding them, forming H II regions. Far ultraviolet photons (FUV, 6 eV < hν < 13.6
eV) dissociate molecular gas beyond the H II region, creating a photodissociation region or
PDR (Hollenbach & Tielens 1999). PDRs are ubiquitous: FUV photons dominate the gas
heating in PDRs and affect the chemistry and physics of gas over a large fraction of the
volume and mass of molecular clouds. The study of interactions between FUV radiation and
molecular cloud gas is therefore important in understanding molecular cloud evolution and
the feedback between massive star formation and subsequent star formation in molecular
clouds.
Observations of PDRs compared with theoretical models probe the physical conditions
in star-forming regions through dust and gas emission mainly in the infrared (IR) wavelength
region of the spectrum. PDRs are the source of most of the IR emission in the Galaxy. Dust
grains and large carbonaceous molecules such as polycyclic aromatic hydrocarbons (PAHs)
absorb radiation from the stars, and re-radiate this energy flux in the infrared, producing
a continuum with solid state and PAH spectral signatures. FUV photons incident on dust
grains and PAHs also cause photoelectric ejection of electrons which then collisionally heat
the gas, a process called the grain photoelectric heating mechanism (e.g., Watson 1972, Bakes
& Tielens 1994, Weingartner & Draine 2001). The gas cools via emission in many IR lines.
Emission from PDRs has been quite successfully modelled (see Hollenbach & Tielens 1999
and references therein). The models are usually parameterized by G0, the ratio of the FUV
flux to the Habing (1968) FUV band flux of 1.6 × 10−3 erg cm−2s−1 characteristic of the
local interstellar radiation field, and n, the density of the gas. For advecting (non-stationary)
models, the flow velocity of the material through the PDR is also a parameter. The FUV
fields in PDRs near OB star-forming regions such as the Orion Bar or M17 SW are typically
-- 3 --
very high, (G0 ∼ 104−5); the PDR gas in these regions is inferred to be very dense with
average densities hni ∼ 104−5 cm−3 (Tielens & Hollenbach 1985b). On larger scales and in
more evolved regions such as the Rosette Nebula, the FUV fields and the average densities
may be significantly lower by factors of 102 − 103.
Infrared, sub-millimeter and millimeter wavelength observations of PDRs indicate that
gas in PDRs is inhomogeneous (van der Werf et al. 1993, Luhman et al. 1998), which is
not surprising since molecular clouds themselves are observed to be highly inhomogeneous
on a wide range of scales from tens of parsecs down to tenths of a parsec (e.g., Genzel 1991).
There is both indirect and direct evidence for clumpiness in PDRs. Spatially extended fine
structure line emission of neutral and singly-ionized carbon near H II regions shows that the
FUV penetrates deeper into the cloud than predicted by homogeneous models, suggesting
that the gas is clumpy in nature (Meixner et al. 1992, Stutzki et al. 1988, Steiman-Cameron
et al. 1997). Further, strengths of high excitation lines of molecular species like CO J=14→13
are observed to be greater than that predicted to arise from gas with density inferred from
other PDR species. It is thus inferred that in regions such as M17 SW and the Orion Bar,
the PDR gas consists of a "low" density (typically ∼ 103−5cm−3) component responsible for
the extended CI and CII emission and high-density clumps (n ∼ 106−7 cm−3) which give
rise to the high excitation CO lines from their surfaces or, more precisely, at a depth of
Av ∼1 from their surfaces (Burton et al. 1990, Meixner et al. 1992, Steiman-Cameron et al.
1997, Hogerheijde et al. 1995). Clumps are also observed directly by mapping PDRs at high
spatial resolution in the 2 µm lines of H2 (e.g., Orion Bar, Luhman et al. 1998; Eagle Nebula,
Allen et al. 1999). The small regions of enhanced emission are interpreted as indicating the
presence of high density clumps (n ∼ 106−7 cm−3) with sizes ≤ 0.02 pc, occupying a very
small volume filling factor ∼ 1 − 2%. However, this interpretation is not very conclusive,
and in this paper we present evidence to show that small clumps with such high densities
may not exist in PDRs. Small differences in temperature within the observed regions can
also explain the observed emission, as H2 emission is highly sensitive to gas temperature
(Marconi et al. 1998).
Dense PDRs like the Orion Bar are spatially thin structures and if these PDRs are
indeed clumpy, the clumps must have sizes smaller than the thickness of the warm PDR,
∼ 0.03 pc for the Orion Bar. The presence of small clumps affects the infrared spectrum from
PDRs by introducing a range of densities in the FUV-illuminated region and by introducing
enhanced advection in clumps that are photoevaporating. One of the goals of this paper is
to investigate the conditions under which small, high-density clumps exist in PDRs.
Small clumps might not exist in significant numbers if the intense FUV field photoe-
vaporates them quickly. The incident FUV field heats up and pressurizes the surface layers
-- 4 --
c ρbcP DR and since mc ∼ ρcr3
of the clumps, causing the layers to expand. This mass loss may cause a complete photoe-
vaporation of the dense clumps, and thus destroy them on short timescales. The mass loss
timescales due to photoevaporation of an FUV-exposed clump can be estimated by assuming
that the heated gas at the surface flows outwards at its sound speed, cP DR. The mass density
at the base of this heated flow, ρb, is lower than the mass density, ρc, in the cold clump, as
we will discuss below. (See Table 1 for a list of frequently used symbols in this paper.) The
mass loss rate is given by dmc/dt ∼ 4πr2
c , the photoevaporation
timescale is tP E ∼ mc/(dmc/dt) ∼ rcρc/(cP DRρb). For a typical clump, and an FUV-heated
surface at 1000K, this is approximately 3 × 103(rc/0.01pc)(ρc/ρb) years. Asymmetric mass
loss from the surfaces of very small clumps can rocket them (e.g., Oort & Spitzer 1955)
out of the PDR back into the cloud. Clumps that survive the rocket effect and remain in
the PDR are likely to lose significant fractions of their mass on short timescales of order
. 104−5 years, depending on the ratio ρc/ρb. The clump evaporation timescale can be longer
or shorter than its residence time in the FUV-illuminated surface region of a GMC. From a
frame of reference moving with the advancing ionization front, interclump material and the
clumps in the GMC are advected into the PDR and ultimately flow into the H II region.
Large clumps survive and they can penetrate the H II region and affect its evolution and
structure, becoming Evaporating Gaseous Globules or "EGGs" (Bertoldi & McKee 1990).
In the present paper, we focus on the heating of an individual clump in a PDR by
FUV photons from a nearby O or B star. We study a range of FUV fluxes incident on a
range of clump sizes and determine the photoevaporative lifetimes of the clumps and the
evolution of their structure. Our investigation is analogous to the study by Bertoldi (1989)
and Bertoldi and McKee (1990) on the effects of EUV heating of clumps in H II regions, and
a generalization of the studies by Johnstone et al. (1998) and Storzer & Hollenbach (1999)
of the evaporation of small clumps and protoplanetary disks ("proplyds", O'Dell et al. 1993)
by EUV and FUV photons.
In a subsequent paper (Gorti & Hollenbach, in preparation) we will discuss the cumula-
tive effect of O and B stars on Giant Molecular Clouds (GMCs), and the relative importance
of EUV and FUV photons in dissociating and destroying GMCs. Ultimately, we are in-
terested in the role of FUV radiation in regulating star-formation in GMCs. Destruction
through photodissociation and photoevaporation of potentially star-forming clumps/cores in
GMCs reduces the star-forming efficiency of GMCs, and the FUV field may thereby play an
important role in regulating or limiting star formation. Photoevaporation of the outer enve-
lope of a collapsing isothermal sphere by FUV radiation could rapidly deplete the material in
the envelope, and may limit the final mass of the star formed. On the other hand, compres-
sion of non-collapsing clumps by shock waves driven by the warm surface gas could possibly
drive the inner cores to instability and gravitational collapse, triggering star formation.
-- 5 --
In §2 of this paper, we discuss the present understanding of GMC structure and existing
models. In §3, we describe the evolution of an H II region and its surrounding PDR as they
propagate into the GMC. Timescales relevant for the FUV heating of a clump in various
contexts are briefly discussed in §4. In §5 we present a simple analytical model which provides
an understanding of the underlying physics of FUV-induced clump evolution. We consider
various possible evolutionary scenarios and their dependences on clump parameters and the
strength of the FUV field. For the more general case, a 1-D numerical hydrodynamics code
has been developed and these results are described (§6). In §7, we estimate the significance
of the rocket effect due to asymmetric mass loss and its effect on clump lifetimes in the PDR,
and in §8 we discuss the implications of our results for observations of PDRs in star-forming
regions. Application to observed PDRs is made in §9. We conclude with a summary of the
present investigation (§10)
-- 6 --
Table 1: List of symbols frequently used in this paper
Symbol Meaning
cc
cP DR
G0
N0
nc0
mc0
rc0
tc
te
tF U V
tP E
ts
vIF
Thermal sound speed in cold clump gas
Sound speed of heated, warm PDR gas
Strength of the FUV field in Habing units, 1.6 × 10−3 erg s−1 cm−2
Fiducial column density equal to 2 × 1021 cm−2
Initial gas number density of clump
Initial mass of clump
Initial radius of clump
Sound crossing time in cold clump gas, rc0/cc
Expansion timescale for clumps with ηc0 > ηcrit
Timescale on which the cold clump gas gets heated to the maximum temperature
in a given FUV field
Photoevaporation timescale
Shock-compression timescale for clumps with ηc0 < ηcrit
Velocity of the Ionization Front moving into the GMC
XP DR Thickness of the PDR in the interclump medium of the GMC, bounded by the IF and DF
α
β
γ
δ0
ηcr
ηc0
ηcrit
ηf
λ
ν
Initial ratio of turbulent to thermal pressures in the cold clump gas
Initial ratio of magnetic to thermal pressures in the cold clump gas
Power law exponent of the variation of density with magnetic pressure
Initial thickness of PDR shell on the surface of a clump (= N0/nc0)
Critical column density for complete photoevaporation of pressure-confined clumps
Normalized initial column density to centre of clump, nc0rc0/N0
Critical column density in a given FUV field towards which unconfined clumps evolve
Final column density of pressure-confined clumps in an FUV field
Photoevaporation parameter, (1 + α + β)(ηc0 − 1)/(2(2ν2 + α)) ; λ = 1 for η = ηcrit
Ratio of sound speeds in warm and cold gas, cP DR/cc
-- 7 --
2. Clumps in Giant Molecular Clouds
Giant Molecular Clouds are observed to have a hierarchical substructure that is highly
inhomogeneous, with density enhancements such as cores, clumps and filaments on all ob-
servable scales, from ∼ 100 pc down to ∼ 0.1 pc. We use the term " massive cores" to denote
large scale density enhancements (& 1 pc in size) and "clumps" to denote smaller scale struc-
tures (∼ few tenths of pc or less in size)1. GMCs and their more massive constituent cores
are strongly self-gravitating, and are supported against collapse by turbulence and magnetic
fields. The internal velocity dispersions of the substructures in molecular clouds are found to
scale with their sizes, with decreasing non-thermal support on the smallest scales (Goodman
et al. 1998). The column density through a GMC or a typical GMC substructure is found
to be approximately a constant N ∼ 1022 cm−2, although significant variations (N ∼ 1021−23
cm−2) about this typical column may occur. (For a recent review on GMCs, see McKee
1999).
The smaller substructures or clumps are not all gravitationally bound and the exact
physical nature of these clumps remains uncertain (Williams et al. 2000). They have been
interpreted as temporary fluctuations in density caused by supersonic turbulence within
the cloud (e.g., Falgarone & Phillips 1990, Scalo 1990). This view is supported by their
apparent fractal nature, which seems to show the same structure on smaller scales down to
a few tenths of a parsec (Elmegreen 1999), and their observed supersonic linewidths (Blitz
& Stark 1986). In this picture, the gravitationally unbound clumps are transient objects
formed by converging supersonic turbulent flows. The non-gravitating clumps are therefore
being constantly formed and destroyed, and have short lifetimes of the order of their sound
crossing time, tc ≈ rc0/cc, where cc is the clump sound speed, and rc0 is the initial clump
radius. The effect of the sudden turn-on of an FUV field on such clumps is to drive a
shock into the clump, producing a smaller, denser clump with a warm evaporative outflow.
Surprisingly, the evaporative timescale is now somewhat longer than tc, as we will show.
Assuming that the FUV flux does not strongly affect the turbulence so that the formation
timescale remains the same, the presence of an FUV field will not greatly change the non-
gravitationally bound clump abundance, but it will compress them and alter their thermal
structure and dynamical evolution.
Another interpretation is that small bound clumps as well as small unbound clumps
1Note that our terminology of cloud substructure is the reverse of what is sometimes used in the literature.
We call the large scale structures " massive cores" and the small scale structures "clumps" to be consistent
with observational work on dense regions in PDRs, wherein the density enhancements are usually called
"clumps". Our notation is also consistent with the notion of "hot cores", where massive stars form.
-- 8 --
represent stable physical entities confined by interclump pressure. Observations through
molecular line spatial maps and velocity channel maps or position-velocity diagrams indicate
the presence of dense coherent structures. Indirect evidence suggests the presence of a low
density (∼ 1 − 2 orders of magnitude less than the clumps) "interclump medium" (ICM)
(see Williams et al. 1995 and references therein). The ICM exhibits high velocity motions
(∆v ≈ 10 km s−1) and therefore may have significant pressure, which could be of thermal,
magnetic or turbulent origin. Clumps in this picture are confined by the ICM, and some of
them may be self-gravitating (Maloney 1988, Bertoldi & McKee 1992). Pressure-confined
clumps live much longer and presumably form more slowly than the gravitationally unbound
clumps of the turbulent model. Therefore, FUV heating of these clumps and their destruction
through photoevaporation will have a much more pronounced effect on lowering the clump
abundance in this interpretation.
3. The Overall Evolution Of A Massive Star Forming Region
Massive stars form in the central regions of clumpy, massive, cores (M & 1000M⊙),
of radius ∼ 1 pc inside GMCs (see review by Evans 1999). These massive cores contain
clumps whose densities may reach nc ∼ 105−7 cm−3, or even higher (Plume et al 1997), and
the clumps are surrounded by interclump gas with lower, but uncertain densities, possibly
nICM ∼ 103−5 cm−3. When a massive star forms, the EUV photons from the star immedi-
ately ionize the nearby interclump gas, and form an H II region. The clumps in the H II
region are also exposed to the ionizing radiation and evaporate (Bertoldi 1989, Bertoldi and
McKee 1990). The FUV photons penetrate further through the interclump gas beyond the
interclump ionization front (IF) and dissociate the molecular gas, forming a dissociation
front (DF) and a PDR.
In the earliest stages of evolution when EUV photons emerge from the forming O/B
star, the IF moves as an R-type front. However, the IF quickly stalls and becomes D-type,
at which point the high pressure H II region drives a ∼ 10 km s−1 shock into the neutral
interclump gas beyond the IF. It is instructive to obtain a sense of sizescale at the R/D type
transition. Assuming the massive star emits φi Lyman continuum photons per second, the
Stromgren radius at this juncture is rs ≈ 0.1φ1/3
e4 pc, where φ49 = φi/1049s−1 and the
electron number density ne4 = ne/104 cm−3. If the central star has an FUV luminosity of
LF U V = 105L⊙ (typical of stars with φ49 ≈ 1), the FUV flux incident on the surrounding PDR
is characterized by G0 ≈ 105. The thickness of the PDR is approximately 1022cm−2/nICM
or 0.3n−1
ICM 4 pc. The DF initially advances ahead of the shock, but as photodissociation
approaches equilibrium with H2 formation, the DF slows down. Eventually, the shock moves
49 n−2/3
-- 9 --
ahead of the DF, and the entire PDR is contained in the post-shock shell. The shock now
impacts cold molecular interclump gas. Hill & Hollenbach (1978) showed that at this stage
the shock velocity has slowed to . 3km s−1, and that, if nICM & 104 cm−3, the shock
completely stalls in t . 105 years. Therefore, there is an initial stage (t . 105 years, for
dense ICM) in the evolution of the clumpy PDRs surrounding H II regions where the clumps
are both shocked and exposed to the photoevaporating effects of the FUV radiation. This
can be followed by a stage where the shock has stalled and unshocked clumps move into the
FUV-illuminated zone.
Two configurations are possible after the shock stalls. If the H II region is still completely
embedded, it becomes stationary with its pressure matched by the molecular cloud pressure
and the ionizing photons completely absorbed by recombining electron/proton pairs in the
H II region. Much more likely is the situation where the H II region breaks through the
surface of the cloud and becomes a "blister" H II region. In this case, the ionized gas can
expand away from the cloud, and into the interstellar medium. The IF then slowly eats its
way into the PDR and the DF similarly advances into the molecular cloud. The velocity of
the IF advancing into the PDR is of order vIF ∼ 1 km s−1, just enough so that the flux of
particles through the IF can balance the heavily attenuated flux of EUV photons reaching
the IF (see Whitworth 1979, Bertoldi & Draine 1996, Storzer & Hollenbach 1998). In such a
case, the clumps entering the PDR may not have been previously shocked and compressed.
Further, in the turbulent GMC model, clumps form and dissipate continually and those
formed in the PDR after the passage of the shock are not affected by it. In both the PDR
around the Trapezium stars in Orion and in the M17 SW PDR (Meixner et al. 1992), there
is no evidence for velocity shifts between PDR gas and the ambient cloud gas, or, in other
words, no evidence for a shock at the present time.
In this paper, we focus only on the
photoevaporation of clumps, whether or not they have been shocked, and do not follow the
potential shredding or flattening of a clump by the passage of a shock wave (see, e.g. Klein
et al. 1994).
The thickness or column depth of a PDR has been defined by Tielens & Hollenbach
(1985) to include all the gas and dust where FUV photons play a significant role in the gas
heating or chemistry. By this definition, in regions like Orion or M17, the PDR extends
to columns of ∼ 1022 cm−2 and includes gas which is cool and molecular (H2, CO), but
where FUV still photodissociates O2 and H2O. We are interested in the surface column N0
of the PDR which is heated to the highest temperatures (Tmax ∼ 100 − 3000K, depending
on G0 and n). For G0/n > 4 × 10−2 cm3, N0 ≈ 1 − 3 × 1021 cm−2 (equivalent to a dust
FUV optical depth of order unity) and is nearly equal to the column where atomic hydrogen
converts to H2 in stationary PDRs (Tielens & Hollenbach 1985). We shall call this region
the "warm PDR" hereafter to distinguish it from the entire PDR. The thickness of the warm
-- 10 --
PDR, N0/nICM , is of order 0.03 − 0.1n−1
ICM 4 pc. The crossing timescale for the IF through
the warm PDR is tP DR ≈ 0.3 − 1 × 105n−1
f 5 years, where vf 5 = vIF /(1 km s−1). Gas
and dust (and clumps) from the opaque molecular cloud interior are therefore advected on
these timescales from the shielded cloud, through the warm PDR, and across the IF into the
H II region. Figure 1 shows a schematic diagram of FUV radiation from nearby O/B stars
impinging on massive cores (large sizescales, e.g., Rosette) and clumps (small scales, e.g.,
Orion).
ICM 4v−1
-- 11 --
HII Region
IF
PDR
DF
Core
OB
star
cluster
xPDR
ICM
ICM
OB
stars
1 pc
HII
Region
IF
Clump
Core
DF
GMC
30 pc
Fig. 1. -- A schematic diagram of PDRs on large scales (e.g. Rosette) and small scales (e.g.
Orion Bar). The figure shows blister H II regions formed by the massive stars, and the PDR
bounded by the ionization and dissociation fronts. Clumps in the PDR are exposed to FUV
radiation and lose mass by photodissociation and/or photoevaporation.
-- 12 --
4.
Impulsive heating versus slow heating
FUV-illuminated clumps are heated at their surfaces on timescales tF U V that could
be either fast or slow compared to the internal sound crossing timescale within the clump,
tc = rc0/cc.
Impulsive heating implies that the FUV heating of the clump surface layer,
characterized by column N0, occurs with tF U V ≪ tc. In this case, the outer region is heated
more quickly than the entire clump can respond, and attains a much higher pressure than the
cold central region. A shock propagates inward and a heated photoevaporative flow expands
off the surface.
Several situations in clumpy molecular clouds may lead to tF U V ≪ tc. For a clump with
rc0 ∼ 1017 cm and cc ∼ 0.3 km s−1 (T ∼ 10K), tc ∼ 105 years. The OB star may "turn-on"
on a timescale < 105 years. Even after this initial flash of FUV flux, the clumps may shadow
each other so that tF U V is given by the time for a clump to move out of shadow, which is
of order rc0/vc, where vc is the clump velocity. In a turbulent, supersonic medium vc may
be larger than cc so that tF U V < tc. Similarly, if clumps are formed in turbulent gas by
supersonic converging flows, the FUV turn-on timescale is roughly the time for the clump to
form, which again is tF U V ≈ rc0/vt, where vt is the supersonic turbulent speed. Therefore,
impulsive heating is likely the most appropriate approximation for the turbulent model of
clumps in molecular gas.
Let us consider the other situation of long-lived clumps confined by the ICM. Here, slow
heating is generally the best approximation, but in certain conditions, impulsive heating
may be more appropriate. A common situation after the turn-on of the OB star may be the
advancement of the ionization front and PDR front into the opaque molecular cloud, which
occurs at speeds vIF ∼ 1 km s−1. In this situation tF U V is the time for the DF and the IF
to move so that the clump is one FUV optical depth closer to the IF, or tF U V ≈ tP DR =
XP DR/vIF , where XP DR = N0/nICM is the thickness of the heated PDR zone in the ICM
of density nICM . The criterion for impulsive heating tF U V < tc, is more easily met by larger
clumps that have longer sound crossing timescales, as tF U V is the same for all clumps. This
gives us a limiting size for clumps likely to be impulsively heated,
rc0 > XP DR
cc
vIF
≈ (0.3 − 1)XP DR,
(1)
which is of the order of the thickness of the interclump PDR region itself. Clumps with radii
smaller than this are slowly heated as they emerge from the opaque molecular cloud into the
FUV-irradiated regions. Since we only consider clumps within PDRs that have rc0 < XP DR,
pressure-confined clumps are generally slowly heated in our analysis.
Clumps that undergo slow heating, tF U V ≫ tc, adjust quasi-statically to the changing
FUV flux. Long-lived, small, pressure-confined clumps advecting into FUV-irradiated regions
-- 13 --
provide a prime example of this slow evolution. Such clumps develop a warm PDR surface
which slowly gets warmer, thicker and less dense. The clump and the heated PDR surface
evolve in near pressure equilibrium, with the density of the surface decreasing with the rise
in its temperature, such that nP DRTP DR = ncTc = pressure of the ICM (assuming thermal
pressure dominates on these small scales). At any given time, the evolutionary state of a
slowly heated clump is the same as the final state of an impulsively heated clump exposed
to the same local FUV flux. However, a gradually heated clump constantly adjusts its
PDR surface to maintain pressure equilibrium as the PDR slowly heats.
In contrast, an
impulsively heated clump undergoes shocks and supersonic expanding flows in its attempt
to attain equilibrium. (We note that there is no final equilibrium state if the interclump
pressure is zero). Rapid photoevaporative mass loss can cause an acceleration due to a
rocket effect in very small clumps and move them back into the shielded cloud. Pressure-
confined small clumps, on the other hand, are not rocketed as they evolve quasi-statically. In
a given FUV field and at a given ICM pressure, the final equilibrium state of a small clump
is either a completely photodissociated, warm (TP DR), and expanded (but with P = PICM )
region; or it may consist of a shielded cold molecular core with the same temperature, density
and pressure as the original clump, but with a heated and more diffuse protective surface
layer of PDR material. In §5.3 we provide analytic solutions for these equilibrium structures.
However, in §5 and §7 we primarily focus on the interesting physical processes which occur
for clumps that are impulsively heated.
5. Analytical models for FUV photoevaporating clumps
We first provide approximate analytic solutions for the time evolution of a clump sud-
denly exposed to FUV radiation, using a few simplifying assumptions. As described in §2, the
nature of the clumps themselves is not well understood, and we consider two simple analytic
models that qualitatively correspond to the two main interpretations. Turbulent unbound
clumps are modelled as constant density structures in vacuum. A clump in vacuum disperses
on a sound crossing timescale, in the absence of FUV illumination. Pressure-confined clumps
are also treated as being of constant density, but have a surrounding ICM with equal pres-
sure. In our models, the ICM only serves to confine the clumps and does not itself evolve or
otherwise affect the clump evolution. We seek solutions for the time evolution of the mass,
size and mass loss rate of a photoevaporating clump.
-- 14 --
5.1.
Initial conditions and assumptions
Clumps are assumed to be dense, small spheres of gas, supported by thermal, turbulent
and magnetic pressures. We assume that the magnetic field B scales with a constant power
of the density, so that the magnetic pressure is
PB = B2/8π ∝ nγ.
(2)
The initial ratio of turbulent and magnetic pressures, PN T and PB, are set by two dimen-
sionless parameters,
α = PN T /PT ;
β = PB/PT
(3)
where PT is the initial thermal gas pressure in the clump. A clump may have significant
turbulent support and observations of clumps in star-forming regions (Jijina et al. 1999)
indicate values of α ranging from ∼ 0 in cold, dark clouds to & 2 in regions of massive
star formation. There is observational evidence to suggest that turbulent support in clumps
decreases on the smaller scales of star-forming clumps (rc0 . 0.1 pc), where α . 1 (Goodman
et al. 1998). Measurements of magnetic fields and hence β and γ are difficult, but present
observational data suggest a wide range of values for β, from ∼ 0 to a few (Crutcher 1999).
The value of γ (∼ 1 − 2) describes how the B field responds to a relatively sudden change
in gas density. For a frozen, uniform magnetic field threading a spherical clump, B ∝ n2/3
or γ = 4/3. For our standard case we adopt values of α = 1, β = 1, and γ = 4/3. We later
discuss the sensitivity of our results to these choices for α, β and γ. As discussed in §4,
for clumps confined by an ICM and advecting into FUV-illuminated interclump zones, the
assumption of tF U V /tc < 1 is equivalent to rc0 & (0.3 − 1)XP DR. Therefore, in this case we
treat only clumps above a certain minimum size, which is of order the size of the thickness
of the heated PDR of the ICM. We do not explicitly include self-gravity in our analytical
model, and only qualitatively estimate the effects of gravity on clump evolution. We treat
gravity more quantitatively in the numerical models.
The FUV radiation field as seen by the clump is spherically asymmetric, and stronger
at the surface facing the source. For simplicity, however, we assume a spherically symmetric
FUV field when calculating the internal dynamical evolution and the thermal structure of the
evaporating clump. Typical albedos of interstellar dust are ∼ 0.5 and some of this scattered
radiation is directed backwards (dependent on the phase factor g, see Henyey & Greenstein
1941), resulting in a backscattered FUV flux on the shielded clump surface about 0.1 − 0.2
times that on the directly illuminated surface (e.g., D´esert et al. 1990, Hurwitz et al. 1991).
Although the posterior intensity is diminished compared to that on the front surface, the
PDR surface sound speed is not a sensitive function of the incident FUV flux G0 (Kaufman
et al. 1999). For field strengths G0 ∼ 103−5 near typical H II regions and densities n & 103−6
-- 15 --
cm−3 characteristic of clumps there, even a back flux 10% of that on the clump surface facing
the source would heat the gas to nearly the same sound speed (cP DR = 3 km s−1 on the
source side, cP DR = 2 km s−1 on the shadowed side). Therefore, the assumption of spherical
symmetry is as good as ignoring the flux of FUV photons on the backside when calculating
the thermal structure and internal dynamics. However, the asymmetry of the flows from the
front and back surface will cause a rocket effect on the clumps and induce motion away from
the source, as first recognized by Oort & Spitzer (1955). We discuss momentum transfer
due to mass loss of the clumps in §7, and estimate sizes of clumps that gain velocities high
enough to move out of the PDR and back into the molecular cloud.
In our analytical models (but not our numerical models) we assume that the transition
from the warm outer PDR layer to the cold molecular core of a clump is very thin, so that
the clump can be modelled as an isothermal core with sound speed cc, surrounded by a
PDR envelope of sound speed cP DR. The gas in the core always remains isothermal in our
analysis, even when shock-compressed to high densities. For our pressure-confined clumps,
it is assumed that the ICM is of low density and does not absorb any FUV photons. In
order to obtain analytic solutions, we do not fully treat transient phenomena and emphasise
quasi-steady state aspects of the flow.
5.2. Analytic model for unconfined clumps
We begin by considering the evolution of a simple configuration: a cloud consisting of
clumps that are spheres of radius rc0 and constant initial density nc0, immersed in a vacuum.
We refer to such clumps hereafter as "turbulent clumps". These are dynamic structures that
in the absence of FUV heating expand at roughly their sound speed, cc. When such a clump is
exposed to an FUV field, the photon flux is attenuated by dust and instantaneously heats an
outer column of N0 ∼ 2 × 1021 cm−2 (thickness δ0 = N0/nc0) to relatively high temperatures
and sound speeds (for example, near H II regions with G0 ∼ 104−5 and nc0 ∼ 105 cm−3,
T ∼ 1000K and cP DR ∼ 3 km s−1).
The evolution of the clump is determined by its initial radial column density from the
centre outwards, nc0rc0, and the ratio ν of sound speeds in the FUV-heated region and the
cold clump material,
ν = cP DR/cc
(4)
The parameter ν can be thought of as a measure of the strength of the FUV field incident
upon the cold clump. We introduce the dimensionless parameter ηc0,
ηc0 = nc0rc0/N0 = rc0/δ0,
(5)
-- 16 --
for the initial column density to the centre of the clump. Note that ηc0 is a measure of the
mean column density through a clump.
For turbulent clumps with ηc0 ≤ 1, the initial clump column density is less than N0, and
the entire clump is immediately heated and photodissociated by the FUV flux. Effectively,
therefore, the FUV flux accelerates the expansion of the cold clump by heating it throughout,
decreasing the expansion timescale by a factor given by the ratio of sound speeds at the two
temperatures, cc/cP DR, or 1/ν. For clumps with ηc0 > 1, the FUV flux heats an outer shell of
gas to a higher temperature, thereby increasing its pressure. If the pressure of the outer PDR
shell is sufficiently high, it drives a shock to the centre of the clump rapidly compressing it,
and the compressed clump proceeds to evaporate on a somewhat longer timescale than the
expansion timescale in the absence of an FUV field. On the other hand, if the pressure in the
heated outer PDR shell is low, the shock does not make it to the centre of the clump. The
shock dissipates, followed by an expansion of the clump, until finally the photoevaporative
flow halts the expansion and proceeds to shrink and compress the gas. There are thus two
distinct evolutionary scenarios for photoevaporating turbulent clumps with ηc0 > 1.
For brevity in the main text, the details of our analysis and the derivations of the results
are presented in the appendices to the paper. We now discuss our solutions and describe the
physics of FUV-heated clump evolution and refer the interested reader to the appendices for
the full analysis.
Clumps exposed to a given FUV field, measured by the parameter ν, are either com-
pressed by shocks or expand to adjust their column density to a critical value, ηcrit. This
critical column density, ηcrit can be derived from conservation of mass and from the condition
of pressure equilibrium at the surface of the cold clump gas (see Appendix A for details),
and is defined by the relation
2(2ν2 + α)
ηcrit − 1
+ β(cid:18)
2
ηcrit − 1(cid:19)γ
= 1 + α + β
(6)
Recall that α and β are the initial ratios of turbulent to thermal pressure and magnetic
to thermal pressure, respectively, and that the magnetic pressure scales as nγ. For typical
values of these parameters, the second term on the LHS is negligible and equation (6) can
be used to define a photoevaporation parameter, λ, as
λ =
(1 + α + β)(ηc0 − 1)
2(2ν2 + α)
.
(7)
For clumps with initial column densities ηc0 = ηcrit, the photoevaporation parameter λ = 1.
Figure 2 shows an η − ν parameter plot with equation (6) depicted for the standard case
with α = 1, β = 1, and γ = 4/3, and for two possible extremes of these parameters. The
-- 17 --
result is not very sensitive to the choice of the parameter γ, and ηcrit only depends on the
sum of α and β. For the standard case and with ηc0 ≫ 1 and ν ≫ 1 which is typical of many
PDRs, equation (6) can be approximated as ηcrit ≈ 4ν2/3. The evolution of FUV-heated
clumps is determined by where they initially lie in the η − ν parameter space.
-- 18 --
Fig. 2. -- Logarithmic parameter plot of the ratio ηc0 of the initial column density through
the clump to the fiducial column N0, and the ratio ν of the sound speed in the heated
PDR gas to that in the cold clump gas, for impulsively heated clumps. The critical column
density ηcrit is plotted against ν for different values of α and β, with γ = 4/3. Also shown
are representative trajectories of column density evolution for clumps with initial column
densities less than, and greater than, ηcrit. Note that, in this case, the abscissa refers to the
instantaneous column density through the clump. At t = 0, ν = 1 and the clump has a
column ηc0. With the turn-on of the FUV field, ν increases rapidly. A clump with ηc0 < ηcrit
is then compressed by a shock to ηcrit and moves right on this parameter plot and then
proceeds to photoevaporate steadily. A clump with ηc0 > ηcrit expands so that its column
density becomes equal to ηcrit and continues to photoevaporate. The initial turn-on of the
FUV field (dotted line) moves the column slightly to the left because a column N0 is quickly
dissociated and lost from the clump. Gravity is ignored in this figure (see Figure 5).
-- 19 --
Clumps with an initial column density ηc0 < ηcrit evolve through a shock-compression
phase, until their dimensionless column density increases to ηcrit (see Appendix B). Clumps
with higher initial column densities expand until their column density decreases to ηcrit (see
Appendix C). The evolution of clumps in two representative cases is qualitatively shown in
Figure 2 through their trajectories in η − ν space. Once the clump column density attains its
critical value ηcrit, further evolution is similar in both cases, and the clump column density
at any later time remains constant at this critical value. We note that this would imply
that observed clumps in a given PDR should typically have the same column density, the
critical value ηcrit given by the ambient FUV field. This is valid for clumps photoevaporating
into a vacuum and only violated in the transient early stages which are typically very short,
of order t ≈ rc0/cP DR. From equation (6), clumps at 10K heated on their surfaces to 100
K (or ν ∼ 3 ) should have columns ηcritN0 ∼ 3 × 1022 cm−2, and for T ∼ 1000K or
ν ∼ 10, ηcritN0 ∼ 3 × 1023 cm−2. As photoevaporation proceeds, and clumps lose mass, they
shrink in size and their average density increases, so that the column through the clump
remains the same, nc(t)rc(t) = ηcritN0.
Shock compression In this case, the initial column density is larger than N0, but smaller
than ηcritN0. The FUV can only penetrate through an outer shell of thickness δ0 = N0/nc0 <
rc0. This shell is heated to a higher temperature and the corresponding increase in pressure
causes the outer shell to expand at cP DR radially outwards at the edge and inwards into
the clump at the inner boundary. Figure 3 shows a schematic diagram of the evolution of a
shock-compressed clump.
-- 20 --
Fig. 3. -- Schematic diagram of evolution of shock-compressed clump. The cold clump is
indicated by the lightly shaded region, the dotted region represents the heated warm PDR
surface of the clump, and the dark shaded region shows shock-compressed gas. Panel c shows
the propagation of the shock front into the clump and panel d depicts the evolution of the
clump after being compressed by the shock into a small dense core.
-- 21 --
As the shell expands, the column density N0 through it is maintained by further pen-
etration of FUV photons, and heating of more cold clump material. The pressure at the
base of the PDR shell is initially higher than the cold clump gas. This pressure difference
drives a shock wave into the clump, which rapidly propagates to the centre. Behind the
shock the cold clump gas is compressed to a pressure approximately equal to the pressure at
the base of the PDR shell. The shock travels inward at roughly cP DR and the entire clump
gets compressed in a time ts ≃ rc(0)/cP DR, where rc(0) = rc0 − δ0 is the initial radius of the
cold clump when the FUV flux is turned on (Fig. 3b). The shock reaches the centre and
compresses the clump to a radius rc(ts) = rs. The shock-compressed gas forms a high-density
core, such that the pressures in the core and at the base of the PDR flow are approximately
equal. Further mass loss from the clump is from the surface of this core. The time evolution
of the mass and radius of the clump for t > ts are given by the following equations for our
standard case, where α = 1, β = 1, and γ = 4/3. For the general solution, refer to Appendix
B.
rc(t > ts) = (cid:18) rs
rc(0)(cid:19)5/4
−
(ηc0 − 1)
ηc0ν (cid:19)!4/5
−
tc
10ηc0ν
dmc
dt
3(ηc0 − 1)2 q9/4(cid:18) t
rc0 (cid:19) 6ν
tc (cid:19)(cid:18)rc(t)
=(cid:18)mc0
qrc(0)(cid:19)9/4
mc(t > ts) = mc(0)(cid:18) rc(t)
ηc0
rc(0)
(8)
(9)
(10)
Here q = (λ/3)1/3, rs ≈ qrc(0)(1 − 3(1 + q)/(ηc0 − 1))4/9 and t > ts = rc(0)/cP DR. The
PDR/core interface propagates inward as the warm gas evaporates and, finally, the entire core
is transformed to heated, photodissociated, expanding PDR material. Using equations (8)
and (10), Figure 4 shows the clump radius and mass as a function of time for a case where
the initial column through the clump is given by ηc0 = 10, and ν = 5 represents the incident
FUV flux. This is a typical column density for clumps in molecular clouds, (see §2) and
for a clump of mass mc, corresponds to a density nc0 = 3 × 106(M⊙/mc)1/2 cm−3. In this
example of ν = 5, the FUV field might heat cold clump gas with initial temperature of ∼30
K to about ∼750 K, or ∼ 10K gas to about 250 K.
-- 22 --
(a)
1
0.8
0.6
0.4
0.2
0
0
(b)
0.5
1
1
0.5
0
0
0.5
1
Fig. 4. -- Time evolution of radius (panel a) and mass (panel b) of cold clump gas for a
clump with ηc0 = 10, ν = 5 that undergoes shock compression. The mass and radius are
given as ratios to the initial mass and radius, and time is in units of sound crossing times
in the initial cold clump. The solid line is the analytical result, showing an initial rapid
decrease in clump radius followed by a slower phase of clump radius evolution. The dashed
line is the result from the 1-D hydrocode and is seen to agree well with the analytical curve.
-- 23 --
Also shown in Figure 4 are the results of a more detailed hydrodynamics code simulation
(described in the next section) for the same parameters. There is very good agreement
between the two results, in spite of the various simplifying assumptions made in arriving
at the analytical solutions. The numerical solution for rc(t) oscillates with time due to the
dynamics of the shock compression which we have neglected in the above analysis. Our
analytic solution assumes a quasi-steady state, whereas the numerical code shows the effect
of overshooting the steady state solution and rebounding.
The clump evolution goes through two distinct phases, which is apparent from the plot
of clump radius with time (Fig.[4a]). The clump initially shrinks rapidly with time, as the
shock compresses the clump in a time ts. The average velocity vb with which the radius
decreases is 0.65cP DR for the chosen parameters. After the entire clump is compressed by
the shock, it decreases in size more slowly, now due entirely to the mass loss from its surface.
As discussed earlier, the column density in the compressed clump remains constant, and the
clump gets denser as its size decreases. In Figure 4b, the mass of the clump is also seen to
decrease with time as the entire clump gets photoevaporated in about a sound crossing time;
again, the analytical results closely follow the result of the numerical hydrodynamical code.
The cold clump mass at t = 0 is less than the initial clump mass because it excludes the
mass contained in the outer column N0 which is instantaneously heated by the FUV flux to
the PDR temperature.
Using the results of the above analysis, we can obtain simple estimates of photoevapo-
ration timescales for a clump exposed to incident FUV radiation. We define the photoevap-
oration timescale, tP E, as the time for the radius of the clump to shrink to zero. The time
tP E, can be easily determined from setting the LHS of equation (8) to zero,
tP E ≈ 0.5η2/3
c0 ν −1/3tc
(11)
or
tP E ≈ 104(cid:16)
nc0
105cm−3(cid:17)2/3(cid:18) rc0
0.01pc(cid:19)5/3(cid:18)0.3kms−1
(cid:19)2/3(cid:18) 3kms−1
cP DR (cid:19)1/3
cc
years
(12)
Larger clumps are longer-lived compared to smaller clumps in a given PDR environment.
Photoevaporation timescales of turbulent clumps in PDRs in typical star-forming regions
are thus of the order of a few clump sound crossing times (see §7). Paradoxically, the sudden
turn-on of FUV radiation on a clump in a vacuum can increase its lifetime, even though the
FUV heats the surface and increases the flow speed from the surface! For turbulent clumps
with lifetimes ∼ tc in the absence of an FUV field, exposure to the FUV field results in
a "pressure-confined" period where the clump is compressed due to pressure of the heated
surface PDR layer. This compression reduces the area of the photoevaporating surface and
can extend the lifetimes of these clumps to a few tc.
-- 24 --
Collapse of clumps driven by shock compression The FUV driven shock wave that
compresses a clump to very high densities, may render it gravitationally unstable to collapse.
Thus star formation may be triggered in a previously stable clump. A simple estimate of
parameters leading to clump collapse can be made without explicitly including self-gravity in
the equations. We use our solutions for the radius of the shock-compressed core (Appendix
B) and compare this with its Jeans length, rJ . If rs < rJ , the clump is stable, otherwise
it undergoes gravitational collapse.
It should be noted here that a clump supported by
magnetic pressure with γ > 4/3 will always be stable to collapse regardless of the external
pressure, as the magnetic pressure increases faster during compression than the gravitational
energy density. Our discussion here is thus applicable for clumps with insignificant magnetic
fields, or with a magnetic equation of state where γ . 4/3. We now solve for the collapse
criterion as a function of initial clump parameters and the FUV field strength as measured by
ν, for an illustrative case where α = 1, β = 1, and γ = 1. The radius of the shock-compressed
core is given by
The Jeans length of the compressed core is given by
rs ≃ rc(0)(cid:18)3(ηc0 − 1)
4(ν2 + 1)(cid:19)1/2
rJ =(cid:18) 3πc2
4GmH ns(cid:19)1/2
c
For collapse rs > rJ which yields
ηc0 > 1 +
3
4(cid:18) πnc0c2
c
GmHN 2
0(cid:19)2/3
(ν2 + 1)−1/3
(13)
(14)
(15)
The factor nc0c2
in PDRs is typically of the order of 106−7 cm−3 K. For nc0c2
becomes
c in equation (15) is the initial thermal pressure in the clump, and for clumps
c = 106 cm−3 K, equation (15)
(16)
ηc0 > 1 +
49
(ν2 + 1)1/3
Figure 5 shows the collapse criterion(Eq.[16]) on the η − ν parameter plot. Clumps with
initial column densities greater than ∼ 40N0 and with initial thermal pressures of 106 cm−3
K cannot support themselves against gravity and are unstable to collapse even in the absence
of an FUV field. This region of instability is to the right of the vertical dotted line in the
plot. Clumps to the left of this line are initially stable, but sufficiently strong FUV fields
may drive shocks that compress them and trigger collapse. The FUV fields near OB stars
may thus trigger star formation in previously stable clumps, and increase the star formation
rate in clouds, in cases where γ < 4/3.
-- 25 --
Fig. 5. -- This figure shows the η − ν parameter space and demarcates the regions where
gravitational collapse can be triggered by shock compression (for γ = 1). The dashed line
indicates the column densities of clumps, with thermal pressure 106 cm−3 K, which are
initially unstable to collapse. Clumps with this pressure and columns to the left of the line
are initially stable. The regions to the right of the solid line are where clumps are driven to
collapse by shocks induced by photoevaporation, for initial thermal pressures in the clump
PT = 106. For higher pressures Pc0, both the dotted (η ∝ P 1/2
c0 ) and solid lines (η ∝ P 2/3
c0 )
move to the right.
-- 26 --
Clumps with an initial expansion phase Clumps with initial column densities ηc0 >
ηcrit as given by equation (6) develop very thin (δ0 ≪ rc0) PDR shells when they are first
exposed to FUV radiation. This very thin shell is initially at a high pressure and drives
a shock into the clump on the inside edge of the shell while the outside edge of the shell
expands at roughly cP DR just as is the case for shock-compressed clumps. However, the
shock stalls before it reaches the clump centre as the PDR pressure rapidly declines due to
the shell expansion. The PDR pressure and clump pressure become equal before the shell
thickness is comparable to the clump radius. We summarize below the detailed evolution of
initially expanding clumps; details appear in Appendix C.
Because of the high PDR sound speed and relatively small thickness, the PDR pressure
initially drops faster than the pressure in the cold clump. The pressure in the cold clump
eventually becomes higher than that in the surrounding expanding PDR shell, and the cold
clump gas expands at its sound speed, cc (as it would in the absence of FUV radiation
since the interclump medium is assumed here to be a vacuum). This expansion takes place
until a time te at which the density (and pressure) in the clump gas drops to that of the
outer PDR layer. For t > te, FUV photons begin penetrating into the formerly shielded
cold clump gas. The warm PDR gas now confines the clump gas, and further expansion of
the cold clump is halted. The cold clump proceeds to lose mass gradually and shrinks due
to photoevaporation, with an evolution similar to the final solution for shock-compressed
clumps. The radius and mass of the clump at times t > te and for α = 1, β ≤ 1, γ ≤ 4/3 are
given by
rc(t > te) = rc(te) −
where rc(te) = rc(0) + ccte,
3cP DR
2ν2 + 1
(t − te)
(17)
(18)
(19)
and
mc(t > te) = mc0
te = tc(cid:18)1 −
1
ηc0(cid:19) (cid:18) 3ν + ηc0 − 1
− 1!
3ν + 2ν2 + 1(cid:19)1/2
rc0 (cid:19)2
(cid:18)rc(t)
(2ν2 + 1)
ηc0
Figure 6 shows the change in radius and mass of a clump with initial parameters ηc0 =
100, ν = 3. The clump expands out initially and loses a small fraction of its mass during
this phase. After its density has dropped, so that there is pressure equilibrium between the
cold clump gas and the heated layer immediately surrounding it, expansion is halted. As
photoevaporation causes mass loss off its surface, the clump shrinks in size, and is completely
heated and photodissociated in about 4 crossing timescales. The numerical results are also
overlaid on the plot, and there is reasonable agreement. The photoevaporation timescale for
-- 27 --
initially expanding clumps can be defined analogously to shock-compressed clumps as (see
Appendix C, Eq.(C7))
tP E ≈
4
3
νtc ≈ 105(cid:18) cP DR
3kms−1(cid:19)(cid:18)0.3kms−1
cc
(cid:19)2(cid:18) rc0
0.01pc(cid:19) years
(20)
-- 28 --
1
0.8
0.6
0.4
0.2
0
1.5
1
0.5
0
0
1
2
3
4
0
1
2
3
4
Fig. 6. -- Time evolution of radius and mass for an initially expanding clump, with ηc0 =
100, ν = 3. Radius and mass are given as ratios of their initial values and time is in sound
crossing time units, tc. The figure shows the analytical result (solid line) and that from the
numerical code (dashed line). The clump radius increases in the beginning as the clump
expands and then shrinks as the clump loses mass due to photoevaporation.
-- 29 --
Therefore in this regime of low ν, the photoevaporation timescale tP E is of order several
sound crossing timescales, and is proportional to ν.
5.3. Analytical model for clumps pressure-confined by an interclump medium
For a cloud model in which clumps are structures in pressure equilibrium with an ICM,
the evolution of FUV-heated clumps is somewhat different. The heated shell or PDR can
now no longer expand indefinitely, but is eventually confined by the pressure of the ICM.
The evolution of the clump is again determined by ηc0 and ν. Clumps with an initial column
density ηc0 < N0, are completely heated and photodissociated instantly. Such clumps expand
until the PDR gas reaches the interclump pressure. Clumps with higher initial column
densities again are only heated on their surfaces to a column N0, and their further evolution
depends on the turn-on time of the FUV field tF U V , relative to tc.
If the clumps are heated impulsively, tF U V ≪ tc, pressure-confined clumps initially
evolve similarly to the turbulent clumps discussed earlier. Clumps with ηc0 < ηcrit are thus
shock-compressed, and mass flows are set up which photoevaporate the clump. However, the
outflow runs up against the interclump pressure and eventually reaches pressure equilibrium
with its surroundings. During this time, the clump may either be completely heated through
and get transformed into a sphere of warm PDR gas, or may only be partially photoevap-
orated. The final configuration for partially photoevaporated clumps consists of a small
remnant cold clump, and an extended warm PDR envelope surrounding it and protecting
it from the FUV flux. The clump, PDR envelope and interclump medium are in pressure
equilibrium. Clumps with ηc0 > ηcrit are not much affected by the FUV field. There is no
expansion of the clump due to the presence of the confining ICM, and the initially heated
thin PDR shell expands to form an expanded, but still thin, protective PDR layer. Con-
tainment of the PDR shell by the ICM prevents further penetration of FUV photons into
the cold clump and there is no additional heating. For these clumps, photoevaporation is
relatively unimportant and they retain large fractions of their initial masses.
Clumps that are heated with tF U V > tc evolve quasi-statically, steadily adjusting them-
selves to the pressure of the heated PDR shell. The evolution of these clumps can be deter-
mined from simple steady-state equilibrium considerations. In their final configuration, the
clumps may retain a fraction of their initial cold gas, surrounded by a warm PDR envelope,
in pressure equilibrium with the ICM. We consider two extremes in density profiles of the
clump, a constant density clump and a truncated isothermal sphere (n ∝ 1/r2) in pressure
equilibrium with the ICM.
-- 30 --
First, we solve for the evolution of a clump of constant density. From conservation of
mass,
nc0rc0
3 = nf rf
3 + nP DR(r3
P DR − r3
f )
(21)
where nf and nP DR are the number densities in the final cold clump core and the PDR, and
rc0, rf , and rP DR are the radii of the initial clump, remnant clump, and the PDR envelope
respectively. As the final configuration is a remnant clump in pressure equilibrium with the
PDR and the ICM, the final density of the core nf = nc0, and
nc0cc
2+αnc0cc
2+βnc0cc
2 = nP DRcP DR
2+αnP DRcc
2+βnc0cc
2(nP DR/nc0)γ (= Pressure in the ICM)
(22)
Further, we know that the column density through the PDR envelope is N0, as the remnant
clump is shielded from heating. Therefore,
Using equations (21) and (23) to eliminate nP DR and rP DR, equation (22) can be written as
nP DR(rP DR − rf ) = N0
(23)
where
ν2 = (1 + α + β)ξ − βξ1−γ − α
ξ =(cid:18)η3
c0 − η3
f −
3
4
η2
f(cid:19)1/2
−
3
2
ηf
(24)
(25)
and ηf = nc0rf /N0. For complete photodissociation ηf = 0, or ξ = η3/2
c0 . For the standard
case α = 1, β = 1 and γ = 4/3, and for ηc0 > 1, ν > 1, equation (24) can be used to define
a critical column density for complete photoevaporation.
ηcr ≃ 0.48ν4/3
(26)
All clumps with an initial column density lower than ηcr are thus eventually completely
photodissociated. Alternately, a clump with an initial column density ηcr has to be exposed
to an FUV field (ν) greater than or equal to that given by equation (26) to be completely
photodissociated.
-- 31 --
Fig. 7. -- The evolution of column density ηf of the clump against the parameter ν. As soon
as the clump is exposed to an FUV field, an outer column N0 is instantaneously heated, and
thus for ν marginally greater than 1, at t = 0, the final column density ηf = ηc0 − 1. The
figure shows how the final column density of the equilibrium clump configuration decreases
with an increasing FUV field. Also shown are the contours depicting the percentage of mass
lost from the clump as it photoevaporates.
-- 32 --
Figure 7 shows a plot of the final column density of a clump, ηf , as it is exposed to
various FUV fields, measured by ν, for α = 1, β = 1 and γ = 4/3. The figure shows
trajectories followed by the clump column density as the FUV field is increased, as might
happen when the clump slowly emerges into the PDR region from the shielded interior of
the molecular cloud. Note that the abscissa is ηf and not the initial column density ηc0.
The outer column of cold clump gas N0 that is initially heated is thereby discounted, and
as soon as the FUV field is "turned-on" at time t = 0, ηf = ηc0 − 1. As the FUV field (ν)
increases, the clump begins to lose mass due to heating of the outer shell, and its column
density begins to decrease with increasing ν. As ν approaches the critical value for complete
photodissociation for that particular ηc0, the column density rapidly decreases with ν and
the clump photoevaporates completely. If the final strength of the FUV field is less than
that required for complete photoevaporation, the clump column density remains fixed at a
point on its trajectory, as determined by the value of the parameter ν. Therefore, given
a clump's initial column density and the strength of the local FUV field, the final column
density of the clump can be determined. Equivalently, the observed ηf and ν can be used
to determine the initial column ηc0 of the clump. The total mass loss of the clump during
its evolution can also be evaluated from equation (24). Figure 7 also shows the contours for
the mass loss being 5%, 50% and 95% of the initial mass, for different values of ηf and ν.
The evolution of a clump with an isothermal density profile can be determined analo-
gously. The initial density distribution in the clump is given by n(r) = ncs0r2
c0/r2, where
ncs0 is the density at the clump surface, and the initial mass of the clump is 4πmHncs0r3
c0.
We assume that the density in the heated PDR shell in the final configuration of the clump
is constant. Utilizing equations( 21-23) for this case, we obtain an equation very analogous
to equation (24) where now
ξ =(cid:18)3η3
cs0 − 3η3
f −
3
4
η2
f(cid:19)1/2
−
3
2
ηf
(27)
ηcs0 = ncs0rc0/N0 and ηf = ncs0rf /N0. An effective average column density through the
clump can be defined by the ratio of the mass to area of the clump, ¯ηc0 = mc0/(πr2
c0mH N0) =
4ηcs0. The evolution of a clump with an isothermal density profile is thus qualitatively similar
to that of a constant density clump, but with a different effective column density for the
same mass. These solutions apply to clumps that are massive enough for gravity to be
important in determining their density structure, and therefore have appreciable central
density concentrations.
-- 33 --
6. Numerical code and results
In the above analysis, simplifying approximations were made, which are now relaxed in
a numerical code to obtain complete and more exact solutions (see Appendix D for details).
One of the major differences between the numerical code and most of the analytic equations
is the inclusion of gravity in the numerical code.
6.1.
Impulsively heated clump in a vacuum
We first discuss the results of a numerical simulation of a clump generated by turbulence
and heated impulsively. This case is hereafter referred to as Case V (for "vacuum"). A
vacuum boundary condition is used so that the clump is free to expand, even in the absence
of any heating. Initially, the thermal, turbulent and magnetic pressures in the clump are all
assumed to be equal (α = β = 1), and the magnetic pressure scales with the density to the
power γ = 4/3. The FUV field is turned on instantly (tF U V =0). The clump is assumed to
be a constant density sphere, with a column density given by ηc0 = 10, and in a FUV field
characterized by ν = 10 in the outer layers. The value of ηc0 corresponds to that typical of
clumps in clouds, for example, mc = 0.8M⊙, nc = 2 × 105cm−3, and rc = 0.03pc. Clumps
exposed to FUV fields with G0 ∼ 105 are heated to TP DR ∼ 1000K on their surfaces; if
their shielded centres are Tc ∼ 10K, then the ratio of sound speeds ν ∼ 10. These initial
conditions imply the clump lies in the shock compression regime of Figure 2, and we expect
from our analytic solutions that it should photoevaporate in less than one sound crossing
timescale tc (from eq. [11])
-- 34 --
10
5
0
1
0
-1
-2
40
30
20
10
0
10
8
6
4
2
0
0
0
0
1
1
1
1
2
2
2
2
R
3
3
3
3
4
4
4
4
Fig. 8. -- Density, velocity, pressure and temperature as a function of radius at times t =
0.05, 0.1, and 0.15tc. The density, pressure and temperature are scaled to the initial values
in the clump and the velocity is in units of the clump thermal speed cc.
-- 35 --
Figure 8 shows the density, velocity, pressure and temperature profiles in the gas as a
function of radius at three different instants of time, t = 0.05tc, 0.1tc and 0.15tc. As the
outer layers heat up and the increased thermal pressure causes them to expand, a shock
propagates into the clump, compressing the cold inner clump gas. The shock strengthens as
it progresses to the centre of the cloud, as evidenced by the increasing density and velocity at
the position of the shock front. The central cold gas gets compressed to very high densities,
mainly due to the convergence of the radially-moving shocked gas. In Figure 8 we only show
the evolution prior to the shock reaching the centre of the cloud. After the shock reaches the
centre, the compressed gas rebounds and the cold inner clump undergoes radial oscillations
as it settles into pressure equilibrium with the warm, expanding PDR outflow in the outer
layers. The mean density of the compressed clump gas increases with time, as expected from
our analytical solutions (eq. [B7]). As the clump loses mass, the FUV penetrates deeper and
the clump is completely photoevaporated in 0.7tc. Our analytical solution (eq. [11]) predicts
a photoevaporation timescale of 0.5tc, which agrees very well with the numerical solution.
The photoevaporation timescale derived for a clump with this column density, ηc0 = 10, and
density, radius and sound speed given by nc = 2 × 105cm−3, rc = 0.03pc, and cc = 0.3kms−1
respectively, is 104 years.
6.2. Gravitational collapse
The compression of a clump by strong shocks due to FUV heating can raise the central
densities by large factors. This, however, did not render the clump gravitationally unstable to
collapse in the above case because γ was chosen to be 4/3, which raises the magnetic pressure
in dense gas sufficiently to prevent gravitational collapse (Chandrasekhar 1961). The scaling
index of the magnetic pressure with density in clumps depends on the orientation of the
magnetic field in the clump, and on whether the field is frozen into the neutral gas. For a
frozen, uniform, unidirectional magnetic field, PB ∝ n4/3. The magnetic field configuration
in clumps is considerably more complicated (Ward-Thompson et al. 2000), and ambipolar
diffusion may allow the neutral gas to slip past magnetic flux lines; making it difficult to
define γ unambiguously.
We investigate the possibility of triggering gravitational collapse in clumps through a
simulation where we adopt γ = 1, and all other parameters identical to those in Case V. Such
a clump is expected to undergo gravitational collapse based on our earlier analysis (eq. [16]
where ηc0 = 10, and Figure 5).
-- 36 --
Fig. 9. -- This figure shows the mass in the clump, core and warm PDR shell as a function
of time. At collapse, the clump mass drops sharply to zero. Also shown in dotted lines are
the clump and PDR shell masses for the case where γ = 4/3.
-- 37 --
The evolution of the clump is initially similar to that in Case V. The outer surface
expands, and a shock is driven into the clump centre. As the shock reaches the centre, the
density of the cold clump gas increases and the clump radius decreases, until the central
region becomes gravitationally unstable. A collapsing "core" is thus formed. This can be
seen in Figure 9, which shows the mass of the clump as a function of time. The mass of the
collapsing core is also shown in the figure, along with the results of Case V. The core mass is
seen to increase above that contained in the clump just before collapse as the core accretes
mass from the warm PDR gas. A significant fraction (60%) of the initial clump mass is driven
gravitationally unstable to collapse, on very short timescales. In the absence of sufficient
magnetic field support, exposure to strong FUV fields can thus trigger star formation in
clumps. We also conducted a simulation with a weaker FUV field (ν = 5), and found that
the central regions collapsed in that case too, but the collapsing core contained only about
10% of the initial clump mass. Although these triggered collapse solutions are interesting
and instructive, we emphasize that for initially stable clumps with β ≈ 1, they hold only for
γ < 4/3, which requires ambipolar diffusion. Since ambipolar diffusion timescales are longer
than the shock compression timescale, it is likely that realistic clumps require γ ≥ 4/3 and
that triggered collapse does not often occur.
6.3. FUV heating of a pressure-confined clump: tF U V ≫ tc
Pressure confined clumps are modelled as Bonnor-Ebert spheres which are in hydrostatic
equilibrium. We construct a Bonnor-Ebert sphere by setting up a density distribution for a
clump with thermal, magnetic and turbulent pressure support and in hydrostatic equilibrium.
The density in the clump decreases as r−2 in the outer regions (similar to an isothermal sphere
distribution), and flattens out in the central regions. The ratio of central to surface density x,
determines the stability of the sphere against gravitational collapse, with a value exceeding
14.4 being "critically unstable" for an isothermal, non-magnetic gas. In these runs with a
confining ICM, the outer 200 zones are used to represent the ICM. The ICM has the same
pressure as the surface of the clump and is a constant density medium with a density (and
temperature) contrast of 1000. As discussed in §4, for clumps confined by an ICM, a slow
turn-on of the FUV field may be more appropriate. The FUV field in the numerical run is
turned on in a time tF U V = 5tc, and we increase the temperature on this timescale so that
the heated outer shell reaches its maximum temperature at 5tc.
Figure 10 shows the results for a case with varying x and where ηc0 = 10, ν = 5, γ = 4/3,
and α = β = 1. For Bonnor-Ebert spheres with x & 6, the outer region with an isothermal
density profile contains more than half the mass, and the clump evolution resembles that of
-- 38 --
an isothermal sphere. In the figure, the mass of the clump is shown with time along with
the analytical results for an isothermal sphere and a constant density sphere with the same
initial column. The numerical results for three values of the central to surface density ratios,
x = 1, 3 and 6 are indicated. For central concentrations greater than that represented by
x = 6, the analytical results for an isothermal pressure-confined sphere match the numerical
results, and the agreement increases as x increases and the Bonnor-Ebert clump configuration
approaches a 1/r2 density profile. For x = 1, the clump is a constant density sphere, and the
corresponding analytical calculations apply. A slight discrepancy can be noted in the figure
between the two results, and this is due to the differences in the more realistic numerical
model, i.e. an exponential drop in the PDR shell temperature with column and the inclusion
of gravity. For intermediate central density concentrations, as represented by the case with
x = 3, the mass loss rate is in between the two analytical extremes, as expected.
-- 39 --
Fig. 10. -- This figure shows the evolution of mass with time for a slowly heated, pressure-
confined clump (Bonnor-Ebert sphere) with varying central to surface density ratios, x = 1, 3
and 6. In this case, ηc0 = 10, ν = 5, and tF U V = 5tc. The corresponding analytical solutions
for a constant density sphere (dotted line) and an isothermal sphere (dashed line) are also
shown for comparison.
-- 40 --
7. Acceleration of clumps due to rocket effect
Spherically asymmetric mass flows from photoevaporating clumps can cause a rocket
effect on the clump, as first noted by Oort & Spitzer (1955) in their theory on cloud acceler-
ation by ionizing radiation from OB stars. Clumps exposed to FUV radiation in PDRs see
a stronger incident flux of photons from the direction of the OB stars, though a significant
flux due to backscattered radiation from dust on the opposite surface tends to reduce this
asymmetry (see §5.2). The front and back hemispheres of the clump (with respect to the OB
star) are thus heated to different temperatures (different ν) and the escaping material has
a higher flow velocity on the star-facing surface of the clump. The net thrust on the clump
accelerates it away from the star, and is known as the rocket effect. With the help of our
analytical expressions for the evolution of an impulsively-heated, photoevaporating clump
in a PDR, we now evaluate a simple criterion which determines the clump size at which the
rocket effect becomes significant.
A simple 1-dimensional formulation for the acceleration due to the rocket effect is given
by
dvR
dt
= −
1
mc
dmc
dt
(vf f − vf b)
(28)
where vf f and vf b are the flow velocities at the front and back surfaces respectively and
are assumed equal to the PDR sound speed cP DR at those surfaces. We first compute the
acceleration due to the flow from each hemisphere separately, take the difference to estimate
the net acceleration, and then calculate the net velocity vR attained by the clump. The mass
loss from a hemisphere is given by 2πrc(t)2ρbcP DR, where ρb is the density at the base of the
flow. Taking only the component of the momentum in the flow along the direction to the
star, we obtain
(cid:18)dvR
dt (cid:19)1/2
=
cP DR
mc
πrc(t)2ρbcP DR
The net acceleration on the clump is therefore given by
(29)
(30)
dvR
dt
=
πr2
f c
mc (cid:2)(ρbc2
P DR)f − (ρbc2
P DR)b(cid:3)
where the suffixes f and b denote the quantities for the front and back surfaces of the clump,
and the clump radius rf c is determined by the dominant flux on the front surface. As
the initial expansion/compression phases in the evolution of a photoevaporating clump are
typically short (see §5), we ignore this initial phase of evolution and assume the clump is
evolving such that its column density is at its critical value ηcrit ∼ 4ν2/3, for the standard
values of α, β and γ (see eq. [6]). Along with the relation for the column through the flow
-- 41 --
nb(t)rc(t) = 2N0 for each surface and the mass of the clump mc = 4πmHnf cr3
appendix, eq. [A2]), we thus obtain for the net acceleration of the clump,
f c/3, (see
dvR
dt
=
9
8
c2
c
rc (cid:20)1 −
(c2
(c2
P DR)b
P DR)f(cid:21)
(31)
The clump will remain in the PDR of the GMC and be subject to photoevaporation if the
rocket velocity vR attained in a time tP DR = XP DR/vIF is less than vIF , or equivalently if
rc >
9
8
c2
c
v2
IF (cid:20)1 −
(c2
(c2
P DR)b
P DR)f(cid:21) XP DR = rR
(32)
There is thus a critical size rR of clumps, where for rc < rR they get rocketed back into the
shielded molecular cloud, and for rc > rR remain in the PDR and get photoevaporated or
survive to photoevaporate in the H II region. For example, for front and back surface flow
velocities of 3 km s−1 and 2 km s−1 respectively, cc = 0.3 km s−1 and vIF = 0.5 km s−1
equation (32) gives a critical clump size rR = 0.225XP DR. Clumps much smaller than this
will get rocketed back into the molecular cloud as soon as they enter the PDR, partially
evaporating in the process. We emphasize again that the rocket effect is significant only for
turbulent clumps which undergo impulsive heating and photoevaporative flows. Pressure-
confined clumps are gradually heated on their exteriors and do not go through phases of
evolution with rapid mass outflows, and hence experience an insignificant rocket effect.
The mass lost by a clump before it gets accelerated out of the PDR can be estimated by
integrating equation (28) with time, where only the components of velocity and mass flow
along the direction towards the source are to be considered. Thus the mass of the clump is
approximately given by
mc(t) = mc0 exp(−2vIF /∆cP DR)
(33)
where ∆cP DR is the difference in the flow velocities on the two sides of the clump. For typical
values of ∆cP DR = 1 km s−1, and vIF = 0.5 km s−1, in the time it takes for vR to reach vIF ,
the clump loses about 63% of its initial mass. Therefore clumps which are rocketed back
into the cold molecular cloud interior also lose significant fractions of their mass before they
attain high enough velocities to keep up with the advancing ionization front.
8. Discussion
We begin our discussion by summarizing our results for both the case where clumps are
produced by turbulence in regions with insignificant interclump pressure and the case where
clumps are initially stable structures confined by interclump pressure.
-- 42 --
Consider first the scenario where clumps are continually generated by turbulence in
the molecular clouds and the PDR. Initially, the evolution of the clump is determined by
If λ > 1, or ηc0 > 4ν2/3 for α = 1, β =
its column and the strength of the FUV flux.
1 and γ = 4/3, there is a very brief period of shock compression but the shock quickly
dissipates before making it to the clump centre and the clump radius shrinks only slightly.
Subsequently, the clump expands until λ ≈ 1, at which point the photoevaporative flow halts
the expansion and proceeds to shrink and compress the clump gas, maintaining λ = 1, or
a column ncrc = 4ν2N0/3. If λ < 1, then the shock propagates to the centre of the clump,
compressing it significantly, followed by an evolution at λ ≃ 1 identical to that above.
During the final λ = 1 evolution of the shrinking, compressing, evaporating clump, the
rocket effect may play a significant role depending on the physical size of the clump at this
time. If rc . rR, the clumps will be accelerated into the molecular cloud, losing a significant
fraction of their mass in the process. If rR < rc . XP DRcc/vIF , the clumps will tend to
evaporate in the PDR region. However, because in the turbulent model clumps are forming
in the PDR, including the region near the IF, some of these clumps will survive to enter an
advancing H II region. If rc & XP DRcc/vIF and an H II region is advancing, then most of
these large clumps will survive to be engulfed by the H II region.
The lifetimes of turbulent clumps are of order several sound crossing times of the initial
size of the clump (Eq.[11] and Eq.[20]). As discussed previously, this is somewhat longer
than clump lifetimes in the absence of an FUV field. Since turbulent clumps are being
continuously formed, the steady-state abundance of clumps could actually increase in the
presence of an FUV field.
Based on these results, the following predictions can be made for the turbulent scenario
of photoevaporating clumps in PDRs.
1. There will be an enhanced population of clumps with columns ncrc ≈ 4ν2N0/3. Note
that this implies that in comparing this population from one PDR to another, the
PDR with the lower ν (e.g., due to lower G0) will have clumps with smaller column
densities.
2. There will be a population of small clumps (rc . rR) moving with velocity v ∼ vIF
into the molecular cloud at Av & 1.
3. Compared with clumps in FUV-shielded regions, the clumps in turbulent PDRs will
be smaller, denser and potentially more numerous.
4. Similarly, the clumps entering the H II regions will be smaller and denser than the
molecular cloud clumps, and will have columns peaked at 4ν2N0/3.
-- 43 --
We note here that in the present analysis, turbulent clumps are geometrically idealized
as constant density spheres. In reality, turbulently generated structures may also be sheet-
like or filamentary in nature. Let us consider two simple representations of these geometries,
disks and cylinders. For these structures to survive immediate photodissociation, the column
density through their shortest dimension has to be greater than N0. Disks with larger column
densities through their midplane establish photoevaporative flows in an FUV radiation field.
When the outer boundary of the heated gas reaches a distance equal to the radius of the
disk, the flow diverges and the density in the flow (refer Eq.[A1] and Eq.[A2]) begins to drop
as r−3 if v ∝ r. Therefore, the mass loss rate from the disk is essentially similar to that of
a spherical clump of size equal to the disk radius (Johnstone et al. 1998) Photoevaporative
flows from a cylindrical structure differ slightly from those off a spherical clump of the same
radius. The density at the base of the flow is now ∝ v−1r−1 ∝ r−2 and therefore, the critical
column density ηcrit ≈ 2ν3/3. The mass loss rate from the cylindrical clump (following shock
compression or initial expansion) no longer depends on the radius and is a constant given by
−2πmH N0cP DRl where l is the length of the cylindrical filament. (cf. Eq.[B11]). In summary,
for turbulent clumps which are likely to be sheets or filaments, the mass loss from a thin
sheet or disk of dimension r is best modelled as a clump of radius r. The mass loss rate from
a disk of radius r and thickness t is similar to that of a sphere of radius r and is ∝ r, but the
evaporation timescale is m/(dm/dt) ∝ r2t/r ∝ rt and hence much smaller than the sphere
(m/(dm/dt) ∝ r2) for t ≪ r. Filaments of radius r evolve somewhat differently than spheres
of radius r; their mass loss rates do not decline as r shrinks. However, the evaporation
timescales evolve similarly for cylinders or spheres since for spheres m/(dm/dt) ∝ r3/r ∝ r2,
whereas for cylinders m/(dm/dt) ∝ r2l/r0l ∝ r2.
Consider next the scenario where clumps are pressure-confined structures which are
bounded by a low-density ICM, and the FUV turn-on timescale is of order the crossing time
XP DR/vIF for clumps through the PDR. For small clumps with rc . XP DRcc/vIF , the FUV
turn-on time tF U V > tc, which means that the FUV quasi-statically heats and expands an
outer PDR surface on the clump, continually maintaining a pressure PICM in the PDR region
and in the cold clump region. Again, the evolution of the clump is determined by its column
and the strength of the FUV flux. If ηc0 < 0.48ν4/3 (for α = 1, β = 1, and γ ≥ 1), the
clump is completely heated and photodissociated as it enters the PDR from the molecular
cloud. If ηc0 > 0.48ν4/3, the clump shrinks to a smaller cold clump surrounded by a PDR
shell. Since the evolution is quasi-static, there is insignificant rocket effect on these clumps.
In this scenario, small cold clumps, protected by PDR shells, can survive passage through
the PDR and into the H II region. However, the minimum initial column for such a clump
is ηc0 ≈ 0.48ν4/3 (for α = 1, β = 1, and γ ≥ 1).
For large clumps, rc & XP DRcc/vIF , tF U V < tc, which means that shocks compress
-- 44 --
the clump and strong photoevaporative outflows are initiated as the clump enters the PDR.
Clumps with 1 < ηc0 < 4ν2/3, are compressed by shocks which propagate to the centre, but
they finally relax to an analogous situation as above, i.e., a thick PDR shell surrounding
a small mass cold clump if ηc0 > 0.48ν4/3 and totally photodissociated otherwise. Clumps
with ηc0 > 4ν2/3 have a brief period of shock compression and PDR expansion, but rapidly
relax to a configuration with a thin PDR shell and most of the initial mass still in the cold
shielded core. All the large clumps experience a rocket effect. However, as shown earlier,
large clumps with rc > rR do not experience enough rocket acceleration to prevent them
from entering the H II region. Therefore, in the ICM confined scenario, a variety of clump
sizes enter the H II region. However, the small clumps entering the H II region were initially
clumps of larger size and mass in the molecular cloud, and they enter the H II region with
large protective PDR shells.
One can compare the abundance of clumps of various sizes in PDRs with the abundance
of clumps in the molecular cloud, under the assumption that ICM-confined clumps are long-
lived (the formation time of clumps exceeds the crossing time XP DR/vIF ). Small cold clumps
are converted to PDRs but larger cold clumps are converted to small cold clumps with PDR
shells and this partially replaces the small clump population. Thus, there will be a significant
drop in the cold clump population with ηc0 < 4ν2/3. However, the population ηc0 > 4ν2/3
is little affected. Therefore, a variety of clump sizes enter the H II region, but there is a
suppression of clumps with columns ηc0 < 4ν2/3 compared to the distribution in the shielded
molecular cloud.
Based on these results, the following predictions can be made for clumps initially con-
fined by interclump pressure.
1. Compared with small cold clumps (ηc0 < 4ν2/3) in similar pressure regions of the
molecular cloud, clumps of the same column density in PDRs will have a smaller
relative population. Correspondingly, small and large clumps enter the advancing H II
regions, but the relative number of small clumps is suppressed.
2. Larger clumps are relatively unaffected, but do experience an initial transient period
of shock compression and photoevaporative flow.
3. Small clumps (ηc0 < 4ν2/3) in the FUV-illuminated region will have small masses
relative to their PDR shells.
We have shown that small clumps tend to be destroyed before they are overtaken by the
ionization front of the advancing H II region. This would affect the number, size distribution,
-- 45 --
and structure of pressure-confined clumps that enter an H II region, and therefore affect the
evolution of the H II region (see Bertoldi 1989 and Bertoldi & McKee 1990 for a discussion
of the effect of photoevaporating clumps on the evolution of an H II region). Clumps that
survive to enter the H II region will thus have undergone significant evolution from their
initial states. These clumps could be considerably denser if they are still vigorously photo-
evaporating as they enter the H II region, where they continue to photoevaporate (Bertoldi
1989, Bertoldi & McKee 1990). We will discuss the propagation of clumps into H II regions
in a separate paper, with application to the Eagle Nebula.
We conclude our discussion by using our results to critique the previous inferences of
dense (∼ 107 cm−3) FUV heated clumps in PDRs and to propose several observational
consequences of our results.
Our photoevaporation models provide arguments against the interpretation that dense
(n ∼ 106−7 cm−3) PDR surfaces of clumps give rise to the observed high excitation CO
lines. The fundamental problem is the propagation of the FUV photons to regions of such
high density. If densities of ∼ 106−7 cm−3 are in fact heated to T ≫ 100K, then the thermal
pressures in these regions are much larger than the pressures in the interclump gas. The only
way this is possible is if: (i) the clumps are gravitationally bound or (ii) the warm surface
regions are photoevaporating, as discussed in this paper. In either case, the density in the
clump will smoothly fall off with radius, n(r) ∝ r−b where b ∼ 2 − 3, until the transition to
the ICM. However, FUV photons which heat the CO are only able to penetrate through a
column N0 ∼ 2 × 1021 cm−2, where
N0 ≈Z ∞
r
n(r)dr ≈
1
b − 1
n(r)r
(34)
This equation reveals the size r of the FUV heated region if a density n(r) is required in this
region. It implies that for the FUV to heat gas at densities & 106 cm−3, the clumps must be
very small with size scales of r ≈ N0/106 cm−3 ≈ 2 × 1015 cm at this density. Gravitationally
bound clumps have densities n ∼ 106−7 cm−3 at sizes ≫ 1015 cm (e.g., Shu 1977). In such
clumps, the FUV would not penetrate to such high densities, but would be absorbed in the
lower density regions further out. Such small scales as 1015 cm also argue against unbound
clumps explaining the high temperature CO emission. We have shown in this paper that
such a small clump will either be rocketed out of the PDR or photoevaporate in roughly a
sound crossing time, or ∼ 300 years, a time so short as to rule out this possibility.
The size scale of 100 AU is suggestive of disks around young stars rather than clumps,
so that one might appeal to photoevaporating disks like the "proplyds" in Orion (e.g., John-
stone et al. 1998). Here, the enormous mass reservoir of the disk greatly lengthens the
photoevaporative timescale. However, the area filling factor of these photoevaporating disks
-- 46 --
required to match the CO high-J line intensities in, for example, the Orion Bar region is of
order ∼ 0.1 (Burton et al. 1990). Such a high area filling factor requires a volume density
of & 3 × 105 stars pc−3, more than an order of magnitude higher than that observed in
the densest region of the stellar cluster near the Trapezium (McCaughrean et al. 1994).
Photoevaporating proplyds are thereby also ruled out.
Therefore, it seems quite unlikely that FUV radiation is heating dense, n ∼ 106−7
cm−3, PDR surfaces of clumps or that photoevaporating proplyds can explain the mid-J CO
emission. We suspect that the CO emission arises in n ∼ 105 cm−3 surfaces of large (& 1016
cm) clumps, and that the chemistry, heating and/or dynamics needs to be modified in PDR
models in order to better match the observations.
There are several observational consequences implicit in the scenario of a strongly pho-
toevaporating (i.e., tF U V /tc < 1) clump. First, the molecular material from the clump
core is effectively advected out into the PDR region, as has been discussed by Bertoldi &
Draine (1996), Storzer & Hollenbach (1998), Hollenbach & Tielens (1999). This can have
strong effects on the chemistry, primarily because H2 can exist in surface regions where it
would have been atomic in a stationary, steady-state case. In turn, the higher abundance
of H2 can chemically enhance the abundance of other minor species, such as H+
3 (Bertoldi
& Draine 1996) and CO and can modify the heating and cooling processes and therefore
the PDR temperature. Initial results (Storzer & Hollenbach 1998 and Storzer 2000, private
communication) suggest that advection will enhance the intensities of the mid-J CO lines,
by creating more warm CO near the surface. Finally, the flow will broaden line widths
to values ∆vF W HM ∼ cP DR. However, it should be noted that many common molecules,
such as CO, HCO+, CS, HCN, CN still exist primarily at column depths N > N0 from
the surface, where the advection velocities are low. Broader linewidths should be seen in
C II(158µm),O I(63µm) and possibly the H2 (2µm) lines.
9. Applications to observed PDRs in well-studied star forming regions
We compare our results with available observational data on clump characteristics from
some well-studied star-forming regions, the Orion Bar, M17SW, NGC 2023 and the PDR
surrounding the Rosette Nebula. Clump column densities (η) can be estimated from mea-
sured or inferred densities and sizes of clumps in these PDRs. Densities are usually inferred
indirectly from comparing observed line intensities and line ratios with available PDR chem-
ical models in most of these cases. Clump sizes are ill-determined, as they are usually too
small to be resolved directly and only upper limits to the sizes are available, with the PDRs
in Rosette and NGC 2023 being possible exceptions. The FUV field strengths in these PDRs
-- 47 --
can be estimated more accurately, as the luminosities of the illuminating O/B stars is rea-
sonably well-determined. We use these field strengths and the dust temperature models of
Hollenbach, Takahashi & Tielens (1991) to estimate the temperature of dust in the shielded
clump interior. As the interior gas temperature is closely coupled to the dust temperature,
we set the two equal and thus determine the sound speed cc in the cold clump gas. The
temperature to which the FUV field heats the surface clump gas is determined from the PDR
models of Kaufman et al. (1999), and thus cP DR is obtained, allowing us to estimate the
value of the parameter ν in each region. Below we first determine η and ν for each observed
region, and then use these parameters, compared to Figures 2 and 7, to infer their nature
and evolution.
Orion Bar The Orion Bar has been observed in many atomic and molecular line transitions
in the sub-millimeter and infrared wavelength regions (e.g. van der Werf et al. 1996, Tauber
et al. 1994, Young Owl et al. 2000, Marconi et al. 1998). The molecular line emission
peaks at about 0.03 pc from the IF, and the gas between this dense ridge and the IF is
mainly neutral atomic hydrogen. The observed line intensities and line ratios appear to be
best explained by a two-component model for the PDR gas, with small (. 0.02 pc) dense
(nc ∼ 106−7 cm−3) clumps, required to match the line emission and an interclump medium
nICM ∼ 104−5 cm−3 which causes the chemical stratification of the edge-on PDR. The local
FUV field is estimated as G0 ∼ 4 × 104. Such an FUV field incident on clumps with these
inferred densities would heat their surfaces to temperatures ∼ 2000−5000 K (Kaufman et al.
1999). The gas temperature in the interior of clumps exposed to this FUV flux is expected
to be ∼ 50 K (Hollenbach et al. 1991). We thus obtain η ∼ 15 − 150 and ν ∼ 8 − 10.
M17 SW Multilevel molecular line observations in CS (Wang et al. 1993, Evans et al.
1987) and in several fine structure lines (C II, Si II, O I, Meixner et al. 1992) indicate that the
M17 SW cloud core consists of numerous high density clumps with densities nc ∼ 1 − 5 × 105
cm−3. Maps of CO+ emission compared with theoretical models suggest that the emission
is produced in the warm surface layers of PDRs in dense clumps (Storzer et al. 1995) of
sizes rc ∼ 0.1 pc. The local FUV field near the H II region/molecular cloud interface is
G0 ∼ 8 × 104, which implies PDR temperatures of about 2000K on the clump surfaces and
cold clump gas temperatures ∼ 60K, therefore, ν ∼ 4 − 9. The column densities of clumps
are typically η = 7 − 35, as estimated from typical densities and sizes (Meixner et al. 1992,
Storzer et al. 1995).
-- 48 --
NGC 2023 The well-studied PDR in the reflection nebula NGC 2023, at a distance of
475 pc, is illuminated by a B star and is well-studied (e.g., Wyrowski et al. 2000, Steiman-
Cameron et al. 1997, Howe et al. 1991). Observations of atomic fine structure lines indicate
the presence of clumpy gas in the PDR with densities ∼ 105 cm−3, and an ICM density of
∼ 103 cm−3 (Steiman- Cameron et al. 1997). The PDR is situated about 0.1-0.2 pc away
from the exciting star, and has a width XP DR = 0.04 pc. The clumps have sizes ∼ 0.05 pc
(Wyrowski et al. 2000), and therefore ηc0 ∼ 5. The FUV field strength has been estimated
at G0 = 104, and from the PDR models of Kaufman et al. (1999), cP DR = 3km s−3. With
a clump gas temperature of ∼ 20 K for this value of G0, we obtain ν = 5. It should be
noted that the level of turbulent energy in the NGC 2023 PDR as estimated from observed
linewidths is low (Wyrowski et al. 2000), which suggests that clumps are not rapidly re-
formed by turbulence in the PDR.
Rosette PDR The Rosette Nebula surrounding the young open cluster NGC 2244, is an
expanding H II region at a distance of 1.6 kpc. The H II region/molecular cloud interface
shows a distinct ridge of emission, with substructure on small scales down to 0.5 pc (Schneider
et al. 1998). The PDR region is very extended, and the FUV photons penetrate deep into
the cloud with an interclump medium density nICM ∼ 10 cm−3 (Blitz 1991, Williams et
al. 1995). The ICM near the edge of the HII region may even by partially ionized gas as
suggested by the pervasive Hα emission observed (Block et al. 1992). This region has a
low FUV field, with G0 ∼ 200 about 15 pc away from the OB star cluster, and dropping to
lower values of 10-50, about 30 pc away from the cluster (Schneider et al. 1998). Clump
densities required to match observed CII line intensities are ∼ 104−5 cm−3. Cold clump gas
temperatures are calculated to range from 10 −15 K. The Rosette PDR thus has ηc0 ∼ 5 −35
and ν ∼ 5, while further away from the H II region, ν ∼ 2 − 3.
Using the above determined values of the parameters η and ν for clumps in the above
PDRs, we can locate them on the parameter plots of Figures 2 and 7. Figure 11 shows an
overlay of our model results with inferred data for both the turbulent and pressure-confined
clump models. The data are consistent with both clump models and, within the limits of the
uncertainties in arriving at "observed" η and ν, agree reasonably well with our predictions.
Turbulent clumps are expected to have columns fairly close to the critical value ηcrit which
is ≈ 4ν2/3. The observed data lie near the η = ηcrit line, indicating that the observed
clumps may be turbulent in origin and undergoing photoevaporation in the FUV field. The
exception is NGC 2023. The clumps in this PDR are observed to have negligible turbulence,
and are thus not expected to have a column equal to ηcrit. The data are also consistent
with clumps being pressure-confined. In such a situation, we would not expect to see clumps
-- 49 --
beyond the 95% mass loss line, as seen in the figure. However, there does appear to be
a weak trend of increasing η with ν, which if real is not easily explained in this model.
The clumps in NGC 2023 are probably pressure-confined, and seem to have lost almost
75% of their original mass through FUV heating. High resolution observations of clumps in
PDRs, with more accurate determinations of the sizes and densities of clumps would help
in locating them more precisely on the η − ν plots, and thus be able to distinguish between
the pressure-confined and turbulent clump models.
-- 50 --
(a)
2
2
2
2
2
2
2
2
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1
1
1
1
1
1
1
1
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
2
1.5
1
0.5
(b)
0
0
0.5
1
1.5
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
1
1
1
1
1
1
1
1
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
2
2
2
2
2
2
2
2
Fig. 11. -- Observational data for some PDRs on the η − ν parameter plot for the turbulent
(panel a) and pressure-confined (panel b) clump models. There is a slight indication that
the columns lie along the ν4/3 lines expected in the turbulent model of clumps, with the
exception of NGC 2023.
-- 51 --
10. Conclusion
We have studied the effects of FUV radiation from young OB stars on the evolution of
dense structures or clumps in photodissociation regions. We find that the ambient FUV field
penetrates through the surface of a dense clump and heats this surface layer to high tem-
peratures, causing mass loss and thereby inducing photoevaporation of the clump. Through
analytic approximations and numerical hydrodynamical calculations, we determined the evo-
lution and lifetimes of clumps, subject to various physical conditions and initial parameters.
The evolution of a clump is mainly determined by three parameters, the ratio of the
initial column density to the column penetrated by the FUV flux, ηc0; the strength of the
FUV field, which we denote by the ratio of the sound speeds in the heated gas and the
cold clump material, ν; and the timescale for the turn-on of the FUV field, relative to the
sound crossing time through the clump, tF U V /tc. We consider the evolution of turbulent,
impulsively-heated clumps and also pressure-confined clumps in a PDR. Impulsively heated
clumps with a photoevaporation parameter λ ≈ 3ηc0/(4ν2) < 1 are initially compressed by
a shock driven by the external high-pressure FUV-heated layer to high densities, and later
lose mass as the FUV gradually penetrates through the entire clump. Clumps with λ > 1
go through an initial expansion phase after the rapid decay of the shock wave. Both clumps
with λ < 1 and λ > 1 evolve toward the λ = 1 condition. The final evolution toward
complete photodissociation and heating occurs at constant column through the clump, the
density increasing as the clump shrinks. Photoevaporation timescales in all these cases are
typically a few tc, suggesting clump destruction times of ∼ 104−5 years, under typical PDR
conditions. Slowly heated clumps evolve quasi-statically, and if confined by an external ICM
pressure, a fraction of the initial cold clump mass may be retained if ηc0 > 0.48ν2. Clumps
with initial column densities ηc0 < 0.48ν2 are completely photoevaporated by the FUV flux.
We predict that in the turbulent scenario, observed clumps should all have columns close
to the critical value for the local FUV field, ηcrit ≈ 4ν2/3. Clumps lifetimes are prolonged by
photoevaporation and we also expect a higher steady-state abundance of clumps in PDRs as
compared to the shielded molecular cloud interior. For pressure-confined clumps, there will
be a decrease in the number of smaller clumps in PDRs, and many clumps will lose substantial
fractions of their mass to extended warm PDR shells around them. We compared our results
to observations of some well-studied PDRs, such as the Orion Bar, M17SW, NGC 2023 and
the Rosette PDR, and find that the data are consistent with both clump models, but perhaps
favour the turbulent clump interpretation. Clumps entering the H II region around an O/B
star with a D-type ionization front are expected to have thus undergone significant evolution
as they pass through the PDR.
-- 52 --
We acknowledge support from the NASA Astrophysical Theory Program under RTOP
344-04-10-02, which supports the Center for Star Formation Studies. U. Gorti was supported
by a National Research Council Associateship. We thank K. R. Bell, M. Kaufman, C. McKee
and H. Storzer for useful discussions.
A. Analytic model: Unconfined clumps with tF U V /tc ≪ 1
We first define several characteristic timescales during evolution. The initial sound
crossing time through the clump is denoted by tc = rc0/cc. We can similarly define a sound
crossing time through the warm PDR layer, as tp = δ0/cP DR. Finally, we define the timescale
for the PDR layer to expand to a thickness comparable to rc0, te0 = rc0/cP DR.
We relate the density nb at the base of the PDR shell (r = rc) to N0 for epochs when
the PDR shell has expanded to a thickness greater than the clump radius (t > te0). The
FUV-heated shell of initial thickness δ0 expands as an isothermal flow into vacuum and the
characteristics of the flow regions are essentially similar to that of the Parker wind solution
with no gravity (Parker 1958). For simplicity, it is assumed here that once the warm PDR
region has expanded to δ & rc or t > te0, the gas leaves the clump surface at cP DR and the
velocity of the heated gas increases linearly with radius. Based on our hydrodynamical code
results, this is a reasonable approximation to within about two clump radii, where most of
the column density in the PDR gas lies. The clump mass loss rate is determined by the
penetration depth of the FUV flux, and hence N0. For a power law density profile in the
shell (n(r) ∝ r−b, b & 2), most of this column arises from the base of the shell, or, more
precisely, from the region between the base (clump surface) rc and ∼ 2rc. From the equation
of continuity and v ∝ r, the steady flow density profile of the warm PDR shell is obtained
as,
n(r) ∝ v−1r−2 ∝ r−3,
(A1)
valid for rc < r < 2rc once the shell has expanded to at least two clump radii (t > te0).
Equation (A1) can be easily integrated over r to obtain the column through the warm PDR
at any given instant of time. Let rc denote the instantaneous cold clump radius and rt the
total radius of cold clump and warm PDR gas. At t = 0, rc = rc(0) = rc0 − δ0. Using the
fact that rt ≫ rc for t > te0, and as the FUV flux always penetrates through a column N0,
we have
n(r, t)dr ≈
nb(t)rc(t) for t > te0
(A2)
N0 =Z rt
rc
1
2
We next write down the equations of momentum flux or pressure. The pressure in the
shocked cold clump gas is the sum of thermal, turbulent and magnetic pressures and given
-- 53 --
by
Pc(t) = nc(t)c2
c + αnc(t)c2
c + βnc0c2
c(nc(t)/nc0)γ
(A3)
where nc(t) is the shocked clump density at time t, nc0 is the initial clump density, α is
the initial ratio of turbulent and thermal pressures, and β is the initial ratio of magnetic to
thermal pressure in the clump. The pressure at the base of the PDR flow, Pb, is given by
Pb(t) = nb(t)(c2
P DR + αc2
c) + βnc0c2
nc0 (cid:19)γ
c(cid:18)nb(t)
+ nb(t)v2
f low,
(A4)
where vf low is the outward flow at the base. We assume that vf low = cP DR, and from
equations (A2) and (A4),
Pb(t) = 2
N0
rc(t)
(2c2
P DR + αc2
c) + βnc0c2
nc0rc(t)(cid:19)γ
c(cid:18) 2N0
(A5)
The boundary between the "shock compression" and "initial expansion" regions of Fig-
ure 2 is determined by comparing equations (A3) and (A5). Values of the parameters ηc0
and ν on this boundary are such that the shock dies out just before making it to the cen-
tre. The pressure in the clump at this instant is about the same as the initial pressure and
rc(te0) ≃ rc(0). In a time t = te0, the outer edge of the PDR shell expands out to 2rc0, and
the initial PDR column begins to decrease appreciably due to spherical divergence, allowing
further FUV flux penetration.
If Pb(te0) > Pc(t = 0), the pressure in the PDR flow is sufficiently high for the shock
wave to reach the centre of the clump, compressing the clump and raising its central pressure
so that Pb(te0) = Pc(te0). Such clumps lie in the "shock compression" of Figure 2. If the
initial parameters are such that Pb(te0) < Pc(t = 0), the shock wave rapidly dies out before
reaching the centre and the clump then stops contracting and begins to expand at a speed
≈ cc, maintaining pressure equilibrium. These two evolutionary sequences are separated by
the (ηc0, ν) boundary obtained setting Pb(te0) = Pc(t = 0), or from equations (A3) and (A5),
2(2ν2 + α)
ηc0 − 1
+ β(cid:18) 2
ηc0 − 1(cid:19)γ
= 1 + α + β.
(A6)
For values of ηc0 and ν typical in PDRs such as Orion (ηc0 ≫ 1, ν2 ≫ 1, α = β = 1, γ =
4/3) this relation can be approximated as ηc0 ≈ 4ν2/3. A photoevaporation parameter
λ = 3(ηc0 − 1)/2(2ν2 + 1) can be defined such that the relation λ = 1 demarcates the two
regions. Clumps with λ < 1 are shock compressed and those with λ > 1 initially expand.
-- 54 --
B. Evolution of shock-compressed clumps
As the flow expands in a shock compressed clump, the clump shrinks by losing mass at
a rate
dmc
dt
= −4πmHnb(t)rc(t)2cP DR
(B1)
Initially, a shock is driven into the clump, but we do not attempt to analytically model the
propagation of the shock wave in detail. We assume that the shock travels inward at the
PDR sound speed and the entire clump gets compressed in a time ts = rc(0)/cP DR, where
rc(0) = rc0 − δ0. We first solve for the clump radius, mass and mass loss rate for t < ts,
during the epoch when the shock propagates to the centre.
We assume that the clump radius decreases with a constant speed vb(. cP DR) so that
The mass of the cold clump decreases with time, and at ts is given by
rc(t) = rc(0) − vbt.
mc(ts) = mc(0) +Z ts
0
(dmc/dt)dt
where mc(0) = 4
3πmHnc0rc
3(0). From equations (A2), (B1), (B2) and (B3), we have
mc(ts) = mc(0) − 8πmHN0rc(0)2(1 − vb/(2cP DR))
(B2)
(B3)
(B4)
We solve for vb and thereby calculate rc(ts). To do this we need to calculate the clump
compression at ts. The clump gets compressed by the shock to a new density n(ts) which
can be determined from the condition of pressure equilibrium at the clump surface. The
momentum flux conservation equation across the "front" which marks the boundary between
clump and FUV heated flow is then
(1 + α)nc(ts)c2
c + βnc0c2
c(nb(ts)/nc0)γ (B5)
c + βnc0c2
c(nc(ts)/nc0)γ = 2nb(ts)c2
P DR + αnb(ts)c2
In moderate to strong FUV fields, the PDR sound speeds and hence the shock velocities
(assumed equal to cP DR) are high, and the resulting compression is also high. As the magnetic
pressure PB, scales with a higher power of the density than the thermal pressure PT , (we
use γ & 4/3) and as the initial magnetic pressure is comparable to the thermal pressure, the
thermal contribution to the pressure in the compressed gas can be ignored. In the expanded
flow, the density is low and here the thermal pressure and dynamical pressure dominate.
Equation (B5) can be simplified to give the density of the compressed gas
nc(ts)
nc0
=(cid:18) (2ν2 + α)
β
nb(ts)
nc0 (cid:19)1/γ
.
(B6)
-- 55 --
From equations (A2) and (B6),
nc(ts)
nc0
=(cid:18)2δ0(2ν2 + α)
β
(cid:19)1/γ
rc(ts)−1/γ.
The mass of the clump at time ts can also be expressed as
mc(ts) =
4π
3
mH nc(ts)rc(ts)3.
From equations (B2), (B7) and (B8)
mc(ts) =
4π
3
mHnc0(cid:18)2δ0(2ν2 + α)
β
(cid:19)1/γ
rc(0)3−1/γ(cid:18)1 −
vb
cP DR(cid:19)3−1/γ
.
(B7)
(B8)
(B9)
At time ts, the radius of the shock-compressed clump is rs = rc(ts) = rc(0)(1 − vb/cP DR)
from equation (B2). The unknown parameter vb is finally determined from equating the two
expressions for the clump mass, equations (B4) and (B9). The evolution of the clump radius
is thus known as a function of time for 0 < t ≤ ts, from equations (B2),(B7) and the solution
for vb. These general equations are, however, fairly complex, and an approximate solution
can be obtained for the case α = 1, β = 1, and γ = 4/3 as,
rs ≈ qrc(0)(cid:18)1 −
3(1 + q)
ηc0 − 1 (cid:19)4/9
(B10)
where q = (λ/3)1/3.
We now solve for the evolution of the clump for t > ts. After being compressed by the
shock, the clump continues to lose mass through photoevaporation, with a mass loss rate,
dmc
dt
= −4πmH nb(t)rc(t)2cP DR = −8πmH N0rc(t)cP DR
(B11)
The mass of the clump at any given time (t > ts) can also be written as
mc(t) = mc(0)(cid:18)2(2ν2 + α)
β(ηc0 − 1)(cid:19)1/γ(cid:18) rc(t)
rc(0)(cid:19)3−1/γ
.
(B12)
From equation (B11) and differentiating equation (B12) with time, the radius of the clump
as a function of time is given
rc(t) = (cid:18) rs)
rc(0)(cid:19)2−1/γ
−
(2γ − 1)
(3γ − 1)
6νηc0
(ηc0 − 1)2 (cid:18) β(ηc0 − 1)
2(2ν2 + α)(cid:19)1/γ(cid:18) t
−
(ηc0 − 1)
ηc0ν (cid:19)!
γ
2γ−1
.
(B13)
tc
-- 56 --
A photoevaporation timescale tP E for the clump can be defined as the time for the radius of
the clump to shrink to zero, and from equation (B13)
tP E = tc (cid:18) rs
rc(0)(cid:19)2−1/γ(cid:18)3γ − 1
2γ − 1(cid:19) (ηc0 − 1)2
6νηc0 (cid:18)2(2ν2 + α)
β(ηc0 − 1)(cid:19)1/γ
+
ηc0 − 1
ηc0ν ! .
(B14)
For α = 1, β = 1, and γ = 4/3, the above equations determining the evolution of the clump
can be simplified as
rc(t) = (cid:18) rs
rc(0)(cid:19)5/4
−
−
(ηc0 − 1)
ηc0ν (cid:19)!4/5
rc(0),
(B15)
tc
10ηc0ν
3(ηc0 − 1)2 q9/4(cid:18) t
mc(t) = mc(0)(cid:18) rc(t)
10qηc0ν (cid:18)1 −
qrc(0)(cid:19)9/4
ηc0 − 1 (cid:19)5/9
3(1 + q)
tP E = tc 3(ηc0 − 1)2
,
+
(ηc0 − 1)
ηc0ν ! .
(B16)
(B17)
and
C. Evolution of clumps that undergo an initial expansion
Clumps with large initial column densities and in low FUV fields quickly evolve to the
point where their internal pressures are greater than that of the (expanded) thin heated
surface layer. They expand out into vacuum (as they would even in the absence of an
external FUV field) at their sound speed cc for a time te, until the pressure drops to that
in the heated outer layer. At t = te, there is pressure equilibrium and equation (B5) again
holds. The clump expands to a new radius rc(te) = rc(0) + ccte and during expansion loses
mass at a rate given by equation (B1). The mass of the clump at t = te is obtained as earlier,
mc(te) = mc(0) − 8πmHN0cP DR(rc(0)te + cct2
e/2).
Also,
mc(te) =
4π
3
mH nc(te)rc
3(te)3.
(C1)
(C2)
Because there is significant expansion, nc(te) ≪ nc0, we assume that thermal pressure dom-
inates in both the clump and the PDR flow. From equations (B5) and (C2)
mc(te) =
8π
3
mHN0
2ν2 + α
1 + α
rc(te)2
(C3)
-- 57 --
Equations (C1) and (C3) can be solved for te, and
te = tc(cid:18)1 −
1
ηc0(cid:19) (cid:18)
3ν + ηc0 − 1
3ν + 2(2ν2 + α)/(1 + α)(cid:19)1/2
− 1!
(C4)
At times t > te, clump expansion is halted, and the clump is now confined by the pressure
at the base of the PDR flow. The clump slowly loses mass, and shrinks to eventually become
completely photoevaporated. The time evolution of clump mass and size can be obtained by
differentiating equation (C3) which also holds for t > te, with respect to time and equating
this with the mass loss rate (equation B11 ). Thus,
rc(t > te) = rc(te) −
3
2
cP DR
1 + α
2ν2 + α
(t − te),
mc(t > te) = mc0
2(2ν2 + α)
rc0 (cid:19)2
ηc0(1 + α) (cid:18)rc(t)
.
(C5)
(C6)
The photoevaporation timescale is obtained by setting the clump radius to zero at tP E to
give
tP E =(cid:18)2ν2 + α
1 + α (cid:19) rc(0) + ccte
3cP DR
+ te,
(C7)
and te is given by equation (C4).
D. Numerical hydrodynamics code
The fluid equations of motion for the system are solved using a 1-D spherical Lagrangian
hydrodynamics code. The equations are solved using a finite difference method, with a
numerical viscosity term, added as a pseudopressure, for accurate handling of shocks and
discontinuities (Richtmyer & Morton 1967, Bowers & Wilson 1991). An isothermal equation
of state is used throughout. The accuracy and stability of the scheme was checked with
standard test problems with known solutions. The computational grid has 1000 equally
spaced radial zones, except for the central 20%, which was spaced logarithmically in radius.
This was done to increase spatial resolution at the centre, and it provides a more accurate
calculation of any shock-induced collapse.
The FUV field is gradually turned on over a timescale, tF U V , which is varied to accom-
modate both the impulsive and slow heating cases described in §4. The temperature of an
outer shell of material is thus raised continuously from Tc to TP DR, reaching its maximum
value of the PDR temperature at a time tF U V . The temperature T , exponentially drops off
with the square of the column, N, into the cloud, T (N) = Tc + (TP DR − Tc)e−(N/N0)2
. This
-- 58 --
profile was chosen to closely mimic the T dependence with column predicted by PDR models
for the values of G0 and n under consideration (Kaufman et al. 1999).
-- 59 --
REFERENCES
Allen, L. E., Burton, M.G., Ryder, S.D. et al. 1999, MNRAS, 304, 98
Bakes, E.L.O., & Tielens, A.G.G.M. 1994, ApJ, 427, 822
Bertoldi F., 1989, ApJ, 346, 735
Bertoldi F., Draine B. 1996, ApJ, 458, 222
Bertoldi F., McKee, C.F. 1990, ApJ, 354, 529
Bertoldi F., McKee, C.F. 1992, ApJ, 395, 140
Blitz, L. & Stark, A. 1986, ApJL, 300, 89
Blitz, L. 1991, The physics of star formation and early stellar evolution, Kluwer, p3
Block, D.L., Dyson, J.E., & Madsen, C. 1992, ApJ, 390, L13
Burton, M.G., Hollenbach, D.J., & Tielens, A.G.G.M. 1990, ApJ, 365, 620
Chandrasekhar, S. 1961, Hydrodynamic and hydromagnetic stability, Oxford:Clarendon
Press
Crutcher, R.M. 1999, ApJ, 520, 706
D´esert, F-X., Boulanger, F., Puget, J.L. 1990, A&A, 237, 215
Weingartner, J.C., & Draine, B.T. 2001, ApJS, 134, 263
Elmegreen, B.G. 1999, The Physics and Chemistry of the Interstellar Medium, p77
Evans, N.J., Davis, J.H., Mundy, L.G., & vandenBout, P. 1987, ApJ, 321, 344
Evans, N.J. 1999, ARA&A, 37, 311
Falgarone, E. & Phillips, T.G. 1990, ApJ, 359, 344
Genzel, R., 1991, Galactic Interstellar Medium, Saas-Fee Advanced Course 21 (Heidelberg:Springer-
Verlag)
Goodman, A.A., Barranco, J.A., Wilner, D.J., & Heyer, M.H. 1998, ApJ, 504, 223
Habing, H.J. 1968, Bull. Astron. Inst. Netherlands, 19, 421
Henyey, L.C. & Greenstein, J.L. 1941, ApJ, 93, 70
Hill, J.K. & Hollenbach, D.J. 1978, ApJ, 225, 390
Hogerheijde, M.R., Janse, D.J., van Dishoeck, E.F. 1995, A&A, 294,792
Hollenbach D.J., Tielens A.G.G.M. 1999, Rev.Mod.Phys, 71,173
Howe, J.E., Jaffe, D.T., Genzel, R., & Stacey, G.J. 1991, ApJ, 373, 158
-- 60 --
Hurwitz, M., Bowyer, S., & Martin, C. 1991, ApJ, 372, 167
Jijina, J., Myers, P.C., & Adams, F.C. 1999, ApJS, 125, 161
Johnstone D., Hollenbach D.J., Bally J. 1998, ApJ, 499, 758
Kaufman M.J., Wolfire M.G., Hollenbach D.J., Luhman M.L. 1999, ApJ, 527, 795
Klein R.I., McKee C.F., Colella P. 1994, ApJ, 420, 213
Kuchar, T.A., & Bania, T.M. 1993, ApJ, 414, 664
Luhman, K.L., Engelbracht, C.W., & Luhman, M.L. 1998, ApJ, 499, 799L
Maloney, P. 1988, ApJ, 334, 761
Marconi, A., Testi, L., Natta, A., Walmsley, C.M. 1998, A&A, 330, 453
McCaughrean, M.J. & Stauffer, J.R. 1994, AJ, 108, 1382
McKee, C.F. 1999, The Origins of Stars and Planetary Systems, eds. C.J. Lada, N.D.Kylafis,
Kluwer, p29
Meixner, M., Haas, M.R., Tielens, A.G.G.M., Erickson, E.F. & Werner, M. 1992, ApJ, 390,
499
O'Dell, C.R., Wen, Z., Hu, X. 1993, ApJ, 410, 696
Oort, J.H., & Spitzer, L.J. 1955, ApJ, 121, 6
Parker, E.N. 1958, ApJ, 128, 664
Plume, R., Jaffe, D.T., Evans, N.J. et al. 1997, ApJ, 476, 730
Scalo, J. 1990, Physical processes in fragmentation and star formation, Kluwer, p151
Schneider, N., Stutzki, J., Winnewisser, G., Block, D. 1998, A&A, 335, 1049
Shu, F.H. 1977, ApJ, 214, 488
Steiman-Cameron, T.Y., Scargle, J.D., Imamura, J.N. & Middleditch, J. 1997, ApJ, 487,
396
Steiman-Cameron, T.Y., Hass, M., Tielens, A.G.G.M., & Burton, M.G. 1997, ApJ, 478, 261
Storzer H., Stutzki, J. & Sternberg, A. 1995, A&A, 296L, 9
Storzer H., Hollenbach D.J. 1999, ApJ, 515, 669
Storzer H., Hollenbach D.J. 1998, ApJ, 495, 853
Stutzki, J., Stacey, G.J., Genzel, R. et al. 1988, ApJ, 332, 379
Tauber, J., Tielens, A.G.G.M., Meixner, M., Goldsmith, P.F. 1994, ApJ, 422, 136
Tielens A.G.G.M., Hollenbach D.J 1985, ApJ, 291, 722
-- 61 --
Tielens A.G.G.M., Hollenbach D.J 1985b, ApJ, 291, 747
van der Werf, P.P., Stutzki, J., Sternberg, A., Krabbe, A., 1996, A&A, 313,633
van der Werf, P., Goss, W.M., Heiles, C., Crutcher, R.M. & Troland, T.H. 1993, ApJ, 411,
247
Wang, Y., Jaffe, D.T., Evans, N.J. et al. 1993, ApJ, 419, 707
Ward-Thomson D., Kirk J.M., Crutcher R., Greaves J.S., Holland w.S., Andre P., 2000,
ApJL, in press
Watson, W.D. 1972, ApJ, 176, 103
Whitworth, A.P. 1979, MNRAS, 186, 59
Williams, J.P., Blitz, L., & Stark, A.A. 1995, ApJ, 451, 252
Williams, J.P., Blitz, L., & McKee, C.F. 2000, Protostars and Planets IV, p97
Wyrowski, F., Walmsley, C.F., Goss, W.M., & Tielens, A.G.G.M. 2000, ApJ, 543, 245
Young Owl, R.C., Meixner, M., Wolfire, M., Tielens, A.G.G.M. 2000, & Tauber, J. ApJ,
540, 886
This preprint was prepared with the AAS LATEX macros v5.0.
This figure "f3.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0204180v1
|
astro-ph/0606307 | 1 | 0606 | 2006-06-13T11:30:30 | Unveiling the nature of the highly absorbed X-ray source SAXJ1748.2-2808 with XMM-Newton | [
"astro-ph"
] | We report on the results of an EPIC XMM-Newton observation of the faint source SAXJ1748.2-2808 and the surrounding field. This source was discovered during the BeppoSAX Galactic center survey performed in 1997-1998. A spatial analysis resulted in the detection of 31 sources within the EPIC field of view. SAXJ1748.2-2808 is clearly resolved into 2 sources in EPIC images with the brighter contributing almost 80% of the 2-10keV flux. Spectral fits to this main source are consistent with an absorbed power-law with a photon index of 1.4+/-0.5 and absorption equivalent to 14E22cm-2 together with an iron line at 6.6+/-0.2 keV with an equivalent width of ~780eV. The significantly better statistics of the XMM-Newton observation, compared with BeppoSAX,allows to exclude a thermal nature for the X-ray emission. A comparison with other observations of SAXJ1748.2-2808 does not reveal any evidence for spectral or intensity long-term variability. Based on these properties we propose that the source is a low-luminosity high-mass X-ray binary located in the Galactic center region. | astro-ph | astro-ph |
Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
Unveiling the nature of the highly absorbed X -- ray source
SAX J1748.2-2808 with XMM-Newton
L. Sidoli1, S. Mereghetti1, F. Favata2, T. Oosterbroek3, and A.N. Parmar2
1 Istituto di Astrofisica Spaziale e Fisica Cosmica -- Sezione di Milano -- IASF/INAF, I-20133 Milano, Italy
2 Research and Scientific Support Department of ESA, ESTEC, Postbus 299, NL-2200 AG Noordwijk, The
Netherlands
3 Science Payload and Advanced Concepts Office, ESA, ESTEC, Postbus 299, NL-2200 AG, Noordwijk, The
Netherlands
Received 12 April 2006; Accepted: 6 June 2006
Abstract. We report on the results of an EPIC XMM-Newton observation of the faint source SAX J1748.2−2808
and the surrounding field. This source was discovered during the BeppoSAX Galactic center survey performed
in 1997-1998. A spatial analysis resulted in the detection of 31 sources within the EPIC field of view. SAX
J1748.2−2808 is clearly resolved into 2 sources in EPIC images with the brighter contributing almost 80% of the
2 -- 10 keV flux. Spectral fits to this main source are consistent with an absorbed power-law with a photon index
of 1.4 ± 0.5 and absorption equivalent to 14+6
−0.2 keV with an
equivalent width of 780+620
−380 eV. The significantly better statistics of the XMM -- Newton observation, compared
with BeppoSAX, allows to exclude a thermal nature for the X -- ray emission. A comparison with other observations
of SAXJ1748.2−2808 does not reveal any evidence for spectral or intensity long-term variability. Based on these
properties we propose that the source is a low-luminosity high-mass X-ray binary located in the Galactic center
region.
× 1022 cm−2 together with an iron line at 6.6+0.2
−4
Key words. Galaxy: center -- X-rays: stars: individual: SAX J1748.2-2808
1. Introduction
SAX J1748.2-2808 is an X -- ray source discovered with the
Narrow Field Instruments on-board BeppoSAX during a
survey of the Galactic Center region (hereafter GC) per-
formed in September 1997 (Sidoli 2000, Sidoli et al. 2001).
The BeppoSAX spectrum was severely absorbed and dis-
played an intense Fe K emission line. The spectrum was
poorly constrained, making both thermal and non-thermal
nature for the X -- ray emission possible. The source, un-
resolved at the angular resolution of the MECS instru-
ment (FWHM ∼1′), is located in the direction of the giant
molecular cloud Sgr D.
of
of
SAX J1748.2-2808 was uncertain and its intense Fe K
line emission, together with its highly absorbed spectrum,
made it a unique object in the GC region which could well
represent the bright tail of a distribution of similar unre-
solved objects significantly contributing to the diffuse Fe
line emission (at 6.7 keV) from the galactic ridge (Koyama
et al. 1989; Ebisawa et al. 2001).
its discovery,
the nature
At
the
time
Interestingly, SAX J1748.2-2808 displays properties
very similar to a class of sources subsequently discovered
Send offprint requests to: L.Sidoli ([email protected])
with the INTEGRAL satellite (see e.g., Walter et al. 2004,
and references therein): these objects show strong pho-
toelectric absorption, hard 2 -- 10 keV spectra, and often
display intense Fe line emission. Most of them also show
X -- ray pulsations, thus indicating that they are likely high
mass X -- ray binaries embedded in a local absorbing gas.
Here we report the results of an XMM -- Newton obser-
vation the region of sky around SAX J1748.2-2808, per-
formed with the main goal of unveiling the nature of this
intriguing source.
2. Observations
The XMM-Newton Observatory (Jansen et al. 2001) in-
cludes three 1500 cm2 X -- ray telescopes each with an
European Photon Imaging Camera (EPIC) at the focus,
composed of and one pn (Struder et al. 2001) two MOS
CCD detectors (Turner et al. 2001). The SAX J1748.2-
2808 field was observed with XMM -- Newton on 2005
February 26-27 for about 50 ks.
Data were processed using version 6.1 of the XMM --
Newton Science Analysis Software (SAS). Known hot, or
flickering, pixels and electronic noise were rejected using
the SAS. A further severe cleaning was necessary because
2
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
Fig. 1. Left panel: Combined (pn+MOS1+MOS2) EPIC image (2 -- 10 keV) centered on SAX J1748.2-2808. Contours
(at 5, 10, 20 and 30 counts/pixel) mark the two sources (the "main" and the "faint") resolved with XMM -- Newton.
Right panel: Optical image of the source field, from the "Second Epoch Survey" of the southern sky made by the
Anglo-Australian Observatory (AAO) with the UK Schmidt Telescope (digitized plates available from STScI at
http://archive.stsci.edu/). The star positionally coincident with the faint source is 0600-28834001 in the USNO-A2.0
catalog.
of the presence of several soft proton flares. After reject-
ing the time intervals where the flares were present, the
net good exposure times reduced to about 32.3 ks for the
MOS1 and the MOS2, and to 13.3 for the pn. Cleaned
MOS1 (with pattern selection from 0 to 12), MOS2 (pat-
tern selection from 0 to 12) and pn (patterns from 0 to 4)
events files were extracted, and used for the subsequent
analysis: a source detection analysis and a spectral anal-
ysis of the brightest sources.
Spectra were rebinned such that at least 20 counts per
bin were present and such that the energy resolution was
not over-sampled by more than a factor 3. Free relative
normalizations between the MOS1, MOS2 and pn instru-
ments were included. The background spectra were ex-
tracted from source free regions of the same observation.
All spectral uncertainties and upper-limits are given at
90% confidence for one interesting parameter.
3. SAX J1748.2-2808: analysis and results
A close-up view of the combined EPIC 2 -- 10 keV image
is shown in Fig. 1, together with an optical image of the
same field. XMM -- Newton resolved SAX J1748.2-2808 into
two sources, a brighter (here called "main" source) and a
fainter one.
In order to minimize contamination from the nearby
fainter source, we extracted source counts centered on the
"main" source, from a circular region of 20′′ radius, for
the MOS1, MOS2 and pn detectors separately.
The fit of the spectrum with an absorbed power-law
(χ2=41.5 with 35 degrees of freedom, d.o.f.) resulted in
structured residuals around 6 -- 7 keV, confirming the pres-
ence of the Fe-K line already observed with BeppoSAX.
Adding a Gaussian line to the power-law model resulted in
a significantly better fit (χ/d.o.f.=24.9/32), with a broad
line (σ=0.43+0.33
−0.20 keV) centered in the range 6.4 -- 6.8 keV
(see Figs. 2 and 3, and Table 1 for the resulting pa-
rameters). We next tried with other simple models, such
as a hot plasma model (mekal in xspec), a thermal
bremsstrahlung spectrum, and a blackbody model (see
Table 1 for the spectral analysis results).
Fig. 2. 68%, 90% and 99% confidence level contours for
the two quantities derived with the absorbed power-law fit
to the "main" source spectrum, absorbing column density
and power-law photon index.
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
3
Table 1. Results obtained fitting the "main" source spectrum. The 2 -- 10 keV flux has not been corrected for the
interstellar absorption. The luminosity, L, has been obtained from the unabsorbed flux, assuming the GC distance
(8.5 kpc). Eline is the Gaussian line centroid in keV, and EW is the associated equivalent width (keV), σ is the line
width in keV, while Iline is the total 10−6 photons cm−2 s−1 in the line.
Column density Parameter
(1022 cm−2)
16.5+6.0
14.3+6.0
−4.5
−4.0
χ2/dof
41.5/35
24.9/32
Flux (2 -- 10 keV)
(10−13 erg cm−2 s−1 )
6.8
6.6
L (2 -- 10 keV)
(1034 erg s−1 )
1.1
1.0
Model
Power law
Power law + line
mekal
Bremsstrahlung
Black body
−2.7
16.6+2.7
16.8+4.8
10.4+3.9
−3.2
−3.2
−0.2
−0.3
−0.5
−0.20
Γ = 1.3+0.6
Γ = 1.4+0.4
Eline=6.6+0.2
σ=0.43+0.33
EW=0.78+0.62
Iline=9.0+7.0
TM = 30+50
Tbr > 12 keV
Tbb = 2.2+0.4
Rbb=0.07+0.02
−0.38
−4.1
−15 keV
38.1/35
41.3/35
−0.3 keV 37.5/35
−0.02 km
6.7
6.8
6.7
1.1
1.1
0.8
Fig. 3. Left panel: Fit to the "main" source spectrum (MOS1, MOS2 and pn data) with a single absorbed power-law
(top panel) together with its residuals in units of standard deviations (middle panel). There is evidence for an excess
around 6 -- 7 keV, indicative of the presence of an emission line. The residuals after adding a Gaussian line to the power-
law model are shown in the bottom panel, in units of standard deviations. Right panel: Photon spectrum resulting from
the best-fit, composed by a power-law together with an emission line at 6.6 keV (see Table 1 for the parameters).
The second fainter source resolved with XMM -- Newton
is too weak to allow a meaningful spectral analysis.
It is reasonable to assume that the SAX J1748.2-2808
emission was mostly contributed by the brighter, "main",
source (and in the following we will call it SAX J1748.2-
2808). The fainter source probably contributed a fraction
of the measured flux from the iron line in the BeppoSAX
spectrum. In order to allow a proper comparison with the
BeppoSAX observation, we extracted a combined XMM --
Newton spectrum from both the main and the secondary
fainter source. The residuals to the fit with an absorbed
power-law again clearly show an excess around 6.5 -- 6.7 keV
(see Fig. 4), requiring the addition of a Gaussian emis-
sion line. The line centroid is 6.6+0.2
−0.1 keV, the normal-
ization is (8+9
−3) ×10−6 photons cm−2 s−1 and the equiv-
alent width is 400+500
−150 eV. The absorbing column den-
sity is (13+7
−3) ×1022 cm−2, while the photon index, Γ,
is 1.2 +0.9
−0.3 (which is harder than the power-law fit to
the "main" source alone, likely due to the hard emission
contributed by the fainter source). The observed flux is
9×10−13 erg cm−2 s−1 (2 -- 10 keV). We then re-extracted
the BeppoSAX MECS spectrum, and deconvolved it with
this best fit model for the XMM -- Newton main plus faint
sources emission. The result is shown in Fig. 5 and demon-
strates that there is no evidence for dramatic changes in
the flux level and spectral shape.
4. A source catalog of the SAX J1748.2-2808 field
Examination of the EPIC images of the SAX J1748.2-2808
field in different energy ranges shows a region rich in faint
X -- ray sources. We performed a detection analysis in order
to obtain a source catalog of this region, using the source
4
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
the final source positions, together with the EPIC com-
bined (MOS1+MOS2+pn) count rates for each source in
each energy range, were calculated using the task emlde-
tect, which performs maximum likelihood fits to the source
spatial count distribution in all energy bands.
This detection procedure resulted in 31 sources, re-
ported in Table 4. Most of them have been detected in
the medium and hard energy ranges, while few sources
were detected only in the soft band (0.5-2 keV), very likely
foreground stars.
The observed fluxes have been calculated from the
combined EPIC count rate, cr, and a total conversion fac-
tor f (flux=f · cr) derived as in Baldi et al. (2002):
Ttot
f
=
TM OS1
fM OS1
+
TM OS2
fM OS2
+
Tpn
fpn
,
where Ttot is the sum of the exposure times from the three
instruments, Ttot=(TM OS1 + TM OS2 + Tpn), and fM OS1,
fM OS2 and fpn are the single count-rate-to-flux conversion
factors.
The sources listed in our catalog show a distribution of
hardness ratios, ranging from sources detected only in the
soft band (likely foreground stars), up to extremely hard
sources, detected only above 5 keV (see Fig. 6). In order
to account for the different source hardness, we derived
three different total conversion factors f , assuming three
kinds of "typical" spectra: for sources detected only in the
"S" band, we considered a blackbody spectrum with tem-
perature 200 eV and a column density of 1021 cm−2 (and
the derived flux is limited to the 0.5 -- 2 keV energy range);
for sources detected only in the "H" band, we have chosen
a power-law spectrum with a photon index Γ=2 and an
absorbing column density of 1024 cm−2 (flux derived in
the 5 -- 10 keV energy range); a power-law spectrum with a
photon index Γ=2 absorbed with NH=2×1023 cm−2 has
been considered for all other sources (flux derived in the
2 -- 10 keV range).
The absorbed fluxes calculated in this way are reported
in Table 4 for the faint sources, while for the brightest
ones, for which a reliable spectrum can be obtained, the
flux measured from the spectral analysis is reported.
A search in the SIMBAD database resulted in two HD
stars positionally coincident with two soft X-ray sources
(see Table 4): HD 316290, located 3.′′1 from source 11, and
HD 161824, at 1′′ from the soft X -- ray source 15. A third
star, Tyc2 929, is located within 1.′′2 from the soft X -- ray
source 28. In Table 3 we report the sky positions and B,
V magnitudes of these 3 optical counterparts, together
with the associated ratios between X -- ray and optical flux
(Maccacaro et al. 1988), which strengthen the physical
association with the soft X -- ray sources.
In the search for counterparts of the X -- ray sources
at other wavelengths, we have conservatively assumed a
circular uncertainty region with a radius of 4′′. The re-
sults from a search in the SIMBAD database are reported
in Table 4 (last column) while a cross-correlation with
the 2MASS All-Sky Catalog of Point Sources (Cutri et
Fig. 4. XMM -- Newton counts spectrum, of the "main"
plus "faint" sources, fitted using an absorbed power-law
(see text for the parameters), together with the residuals
in units of standard deviations
Fig. 5. BeppoSAX MECS2+MECS3 spectrum (data
points) comprising the emission of both the "main" and
"faint" sources, unresolved by this instrument. The line is
the XMM-Newton best fit model to the sum of the two
sources (Fig. 4) folded through the BeppoSAX response.
detection procedure described in Baldi et al. (2002). All
the source detection chain uses SAS version 6.1 tasks.
The cleaned events files were used to produce MOS1,
MOS2 and pn images and exposure maps (which also in-
clude the vignetting effect) in 4 energy ranges: 0.5 -- 2 keV
(soft band, hereafter "S"), 2 -- 5 keV (medium band, "M"),
5 -- 10 keV (hard band, "H"), and 0.5 -- 10 keV (total band).
For each energy band independently, the MOS1, MOS2
and pn images (and the corresponding instrumental expo-
sure maps) were then merged in order to get a higher
signal-to-noise ratio in the source detection. After the
production of a detector mask (with the task emask), a
source detection in local mode was performed in each en-
ergy band separately with the task eboxdetect to produce
a preliminary list of sources using a sliding box technique.
Then, with task esplinemap, all the sources in the list were
removed from the image and the resulting source -- free im-
age was fitted with a cubic spline function in order to
create a background map for each energy band. Then, a
second run with eboxdetect in map mode was made, this
time using the background maps produced before. Lastly,
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
5
Fig. 6. Hardness ratio versus estimated flux in the energy range 2 -- 10 keV for the X -- ray sources in the SAX J1748.2-
2808 field. On the left the hardness ratio HR1 ((M -- S)/(M+S)) versus flux is shown, while in the right panel the
distribution of the hardness ratio HR2 ((H -- M)/(H+M)) versus observed flux is reported. Fluxes are in units of
10−13 erg cm−2 s−1 (see Table 4).
Table 3. Stars positionally coincident with 3 soft X -- ray sources in our catalogue.
Star
R.A. (J2000) Dec. (J2000) B
HD 316290
HD 161824
Tyc2 929
267.203000
267.212250
267.045025
−28.016972
−28.246139
−28.308665
(mag)
10.25
9.63
12.27
V
(mag)
9.76
8.33
10.55
Sp. type
Associated
X -- ray source ID
log(fX /fV )
F8
K1/K2III
11
15
28
−4.3
−4.5
−4.8
Table 2. X -- ray spectral results for the brightest sources
detected in the SAX J1748.2-2808 field (the source ID
is the same as in Table 4). The flux has not been cor-
rected for interstellar absorption and is in units of 10−13
erg cm−2 s−1 in the energy range 2 -- 10 keV, except for
source 6, which is in the 0.5 -- 2 keV range. Al the source
spectra have been fit with a power-law, except for source 6,
where a blackbody has been used (the blackbody temper-
ature, Tbb, is in keV).
Src ID NH
Parameter
1
2
4
5
6
7
−2
−0.3
(1022 cm−2)
1.3+0.3
8+2
15+5
0.7+0.8
< 0.2
1.6+1.3
−0.4
−6
−0.7
−0.1
−0.6
Γ = 1.7+0.2
Γ = 2.0+0.4
Γ = 1.6+0.9
Γ = 1.8+0.8
Tbb=0.2 ± 0.1
Γ = 2.1+1.2
−0.9
−0.6
−0.6
Flux
(2 -- 10 keV)
5
20
30
2
0.5 (0.5 -- 2 keV)
1.4
al. 2003) resulted into 68% X -- ray matches with 2MASS
counterparts.
Within the 14′ radius of the X -- ray field of view,
the 2MASS catalog (Cutri et al. 2003) list 26,413
stars, translating into a surface density µ ∼ 1.19 ×
10−2 sources arcsec−2. This corresponds to 0.6 sources
within each error region of 4′′ radius. Therefore it is
likely that a large number of the 2MASS counterparts are
just random coincidences, as suggested also by the fact
that several X -- ray sources are positionally associated with
more than one infrared counterpart. At the spatial reso-
lution of XMM-Newton, stellar confusion in the Galactic
plane prevents from unambiguously associating infrared
sources with the X -- ray ones. Thus, we will not discuss
further the possible association with 2MASS sources. On
the other hand, for the two brightest HD stars, we estimate
a probability of chance coincidence around 0.02%. This,
together with the measured log(fX /fV ), confirm the real
association of the brightest stars with the X -- ray sources.
Among the brightest X -- ray sources, for which a spec-
tral analysis has been possible, few can be firmly identified
with known objects. Source 2 in Table 4 is a transient
discovered with EPIC in an XMM-Newton observation
performed on 12 March 2003, pointed on the composite
SNR G0.9+0.1 (Sidoli & Mereghetti 2003; Sidoli et al.
2004). The observed flux during the discovery observation
was 3.7×10−12 erg cm−2 s−1 (2 -- 10 keV), almost a factor
of 2 higher than that measured in February 2005, while
the spectral parameters remained constant, within the un-
certainties. The detection of the transient in 2005 could
indicate that we are observing a second outburst from
the source, or that the source is still in outburst since
2003. Source 4 is the pulsar wind nebula in the supernova
remnant G0.9+0.1 (Mereghetti et al. 1998). The spectral
parameters are similar to those measured during previ-
6
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
ous observations with XMM -- Newton (Porquet et al. 2003;
Sidoli et al. 2004). Source 6 is positionally coincident with
source 80 in the ROSAT catalog of the GC sources (Sidoli
et al. 2001b). Compared with the ROSAT observation, it
displays variability.
Five faint X -- ray sources have been detected only above
5 keV. While it is somehow expected that the faintest
sources be more distant and absorbed, on the other hand
it is remarkable that they are observed only above 5 keV,
meaning that their spectrum is both truly hard and ab-
sorbed. Few of these faint hard sources could be back-
ground AGNs. From the Log(N) -- Log(S) measured in the
energy range 5 -- 10 keV (Baldi et al. 2002) with XMM --
Newton, the expected number of hard sources with flux
larger than 5×10−14 erg cm−2 s−1 is 5 -- 11 sources deg−2,
which translates into 1 -- 2 extragalactic hard X -- ray sources
in the XMM -- Newton field of view. The remaining sources
are probably CVs located close to the GC distance (their
fluxes translate into luminosities in the range ∼1032 --
1033 erg s−1).
5. Discussion and Conclusions
In Sidoli et al. (2001) we reported the discovery of a
new X -- ray source in the direction of the Sgr D region,
SAX J1748.2-2808. Our new XMM -- Newton observation
allows to resolve it into two sources (sources 3 and 12 in
Table 4), with a brighter "main" source contributing al-
most 80% of the source flux in the 2 -- 10 keV energy range.
The fainter source is harder (detected only above
5 keV) than the "main" one. A possible optical counter-
part is the star 0600-28834001 of the USNO-A2.0 catalog
(B=18.1, R=13.4), which is listed as [RHI84]10-672 in the
Raharto et al. (1984) catalog of M-type stars. The deriva-
tion of log(fX /fV ) is highly uncertain, but assuming, e.g.,
a blackbody emission at kT∼1 keV, absorbed with a col-
umn density of 1024 cm−2, the 5 -- 10 keV flux translates
into a 0.3-3.5 keV flux ∼4×10−11 erg cm−2s−1 (corrected
for the absorption), and to a log(fX /fV )∼1.8, clearly not
stellar. Thus, the hardness of the X -- ray emission excludes
a coronal emission for the fainter source.
The refined sky position of the brighter source allows
to reject all the possible associations discussed in Sidoli
et al. (2001). The BeppoSAX spectrum was affected by
a high interstellar absorption, NH∼1023 cm−2, suggesting
that the source is probably located at the GC distance
(in this case the luminosity in the 2 -- 10 keV energy band
is ∼1034 erg s−1). A strong Fe K line was present (with
a line centroid of 6.62 ± 0.30 keV), and a good fit was
obtained both with a power-law plus a Gaussian line, and
with a hot thermal plasma model with a temperature, kT,
of 6+35
−4 keV. Thus, the BeppoSAX spectrum was consis-
tent with both thermal and non-thermal models.
The significantly better statistics of the XMM -- Newton
spectrum and the smaller uncertainties in the spectral
slope, favor a non-thermal nature for the X -- ray emission
of the "main" source. A hard power-law (Γ ∼ 1.4) is a good
fit to the data, with an iron line and a high photoelectric
absorption (NH= 10 -- 20 ×1022 cm−2). The absorption is
probably not intrinsic, since the source is located within
about 1 degree from the direction of the Galactic center.
Thermal models do not fit the X -- ray spectrum as well,
and result in very high temperatures (for example, a ther-
mal plasma should be hotter than 15 keV). Among the
thermal models tried, the blackbody is the best in fitting
the spectrum, but results in a high temperature (∼2 keV)
and in an emitting region of less then 0.1 km at the galac-
tic center distance. Thus, the X -- ray spectral shape favors
a non-thermal nature for the X -- ray emission.
The X -- ray emission appears to be stable; the "main"
source has been detected at large off-axis angle during
two previous XMM -- Newton observations performed in
September 2000 and March 2003 (both pointed at the SNR
G0.9+0.1). In both occasions, SAX J1748.2-2808 did not
show evidence for any strong flux variability. Moreover,
the total emission from "main" plus "faint" source, is com-
patible with that observed with BeppoSAX in 1997 (see
Fig. 5).
These properties are suggestive of three possibilities: a
binary system containing a compact object, a background
AGN, or reflection from a molecular cloud core (e.g., sim-
ilar to the X -- rays emission and fluorescent iron line pro-
duced from the molecular cloud Sgr B2; Revnivtsev et
al. 2004). This third possibility, already discussed in Sidoli
et al. 2001, seems now unlikely, based on the high spatial
resolution of the XMM -- Newton observation. The compact
cores contained in the giant molecular cloud Sgr B2, for
example, are about 1 pc in size (Lis & Goldsmith 1991),
while the XMM -- Newton spatial resolution (FWHM∼6′′)
allows us to exclude a source with a size larger than
∼0.25 pc at a distance of 8.5 kpc.
The shape of the X -- ray spectrum and the parame-
ters of the iron line are consistent with a background
AGN. It should be a nearby object, since the Fe line is
not red-shifted. Assuming a distance of 5 Mpc, the 2 --
10 keV unabsorbed flux corresponds to a luminosity of
∼3.4×1039 erg s−1, which is quite low, but still compat-
ible with a low-luminosity Seyfert galaxy (Terashima et
al. 2002). Note that no radio counterpart is present in
the NED catalogue within 30′′ of the X -- ray position, and
SAX J1748.2-2808 does not show evidence for X -- ray vari-
ability on timescales of years, while X -- ray temporal vari-
ability and presence of radio emission are typical proper-
ties of AGNs.
The X -- ray spectral properties of SAX J1748.2-2808
are reminiscent of the soft gamma-ray sources discovered
with the INTEGRAL satellite (see e.g., Kuulkers 2005 for
a review). Several of their X -- ray counterparts display hard
and heavily absorbed spectra, together with intense fluo-
rescent Fe line emission, indicative of dense gaseous en-
velopes around the compact object, illuminated by the
central source. In few of them, the association with OB
optical counterparts and the detection of X -- ray pulsa-
tions, suggest that they are highly absorbed HMXRBs,
not detected in previous surveys at soft X -- rays. The de-
rived luminosity of these INTEGRAL sources is around
L. Sidoli et al.: Unveiling the nature of SAX J1748.2-2808 with XMM-Newton
7
Ebisawa, K., Maeda, Y., Kaneda, H., et al. 2001, Science, 293,
1633
Hog, E., Fabricius, C., Makarov, V.V., et al. 2000, A&A, 355,
L27
Jansen, F., Lumb, D., Altieri, B., et al. 2001, A&A, 365, L1
Koyama, K., Awaki, H., Kunieda, H., et al. 1989, Nature 339,
60
Kong, A.K.H., McClintock, J.E., Garcia, M.R., et al. 2002,
ApJ, 570, 277
Kuulkers, E., in "Interacting Binaries: Accretion, Evolution
and Outcomes", Eds. L.A. Antonelli, et al., Proc. of the
Interacting Binaries Meeting of Cefalu, Italy, July 2004,
AIP, in press, astro-ph/0504625
Lis, D. C., Goldsmith, P. F. 1991, ApJ 286, 232
Maccacaro, T., Gioia, M.I., Wolter, A., et al. 1988, ApJ 326,
680
Masetti, N., Dal Fiume, D., Amati, L., et al. 2004, A&A, 423,
311
Mereghetti, S., Sidoli, L., Israel, G. L., 1998, A&A, 331, L77
Pfahl, E., Rappaport, S., Podsiadlowski, Ph., 2002, ApJ, 571,
L37
Porquet, D., Decourchelle, A., Warwick, R. S., 2003, A&A, 401,
197
Raharto, M., Hamajima, K., Ichikawa, T., et al. 1984, Ann.
Tokyo Astron. Obs., 19, 469
Revnivtsev, M.G., Churazov, E.M., Sazonov, S.Yu., et al. 2004,
A&A, 425, L49
Sidoli, L., 2000, PhD Thesis, Univ. of Milan, Italy
Sidoli, L., Mereghetti, S., Treves, A., et al. 2001, A&A, 372,
651
Sidoli, L., Belloni, T., Mereghetti, S., 2001, A&A, 368, 835
Sidoli, L., Mereghetti, S., 2003, ATel 147
Sidoli, L., Bocchino, F., Mereghetti, S., et al. 2004, MmSAI,
75, 507
Struder, L., Briel, U., Dennerl, K., et al. 2001, A&A, 365, L18
Suchkov, A.A., Hanish, R.J., 2004, ApJ, 612, 437
Terashima, Y., Iyomoto, N., Ho, L.C., et al. 2002, ApJ Suppl,
139, 1
Turner, M. J. L., Abbey, A., Arnaud, M., et al. 2001, A&A,
365, L27
Verbunt, F., Lewin, W.H.G., 2005, to appear in "Compact
Stellar X -- ray Sources", eds. W.H.G. Lewin and M. van
der Klis, Cambridge University Press, astro-ph/0404136
Walter, R., Courvoisier, T. J.-L., Foschini, L., et al. 2004,
Proc. of the V INTEGRAL Workshop, 16-20 February
2004, Munich, Germany. V. Schoenfelder, G. Lichti, &
C. Winkler. ESA SP-552, Noordwijk: ESA Publication
Division, ISBN 92-9092-863-8, p. 417-422.
Wang, Q.D., Gotthelf, E.V., Lang, C.C., 2002 Nature, 415, 148
1036 erg s−1, although there is a large uncertainty in the
distance estimates, and the true luminosity could be much
less than this. SAX J1748.2-2808 lies in the direction of
SgrD molecular cloud, near to SgrB2, which is an im-
portant site of star formation, so it is not unlikely that
SAX J1748.2-2808 is indeed a HMXRB. The low X -- ray
luminosity suggests that it belongs to a class of massive
X -- ray binaries with low persistent emission (in the range
1034 -- 1035 erg s−1), wind-accreting and with no outbursts
(e.g., 4U 2206+54, Masetti et al. 2004). On the other hand,
these sources typically show temporal variability on differ-
ent timescales (sometimes with flares), which has not been
observed in SAX J1748.2-2808 (perhaps because of the
poor coverage). However, wind-fed HMXRBs are usually
quite stable X -- ray emitters on long timescales (months or
years). Low luminosity wind-accreting neutron stars has
been predicted by Pfahl et al. (2002), who proposed that
most of the faint sources detected in the Chandra survey
of the GC (Wang et al. 2002) could be of this kind. A
search for hard unidentified sources from ROSAT PSPC
observations seems to confirm that a new class of fainter
wind-fed X -- ray binaries exists in our Galaxy (Suchkov &
Hanisch 2004).
Other kinds of galactic X -- ray binaries, containing neu-
tron stars or black-holes, seem to be unlikely; the lumi-
nosity (∼1034 erg s−1) suggests an object in quiescence:
but low-mass X-ray binaries (LMXRBs) in quiescence
(soft X -- ray transients) typically have much softer spectra
(blackbody temperatures ∼0.1 -- 0.3 keV; e.g., Verbunt &
Lewin 2005), while black-hole X -- ray novae in quiescence
have much lower luminosities (Kong et al. 2002).
In conclusion, among the different hypotheses dis-
cussed above, the spectral shape (hard, non -- thermal), X --
ray luminosity, the presence of Fe line emission, seem to
favor a low luminosity HMXRB.
Acknowledgements. Based on observations obtained with
XMM-Newton, an ESA science mission with instruments
and contributions directly funded by ESA member states
and the USA (NASA). We thank Giovanna Giardino, Nicola
La Palombara and Silvano Molendi for useful discussions. This
research has made use of the SIMBAD database, operated
at CDS, Strasbourg, France. This publication makes use of
data products from the Two Micron All Sky Survey, which
is a joint project of the University of Massachusetts and the
Infrared Processing and Analysis Center/California Institute
of Technology, funded by the National Aeronautics and Space
Administration and the National Science Foundation. The
XMM -- Newton data analysis is supported by the Italian Space
Agency (ASI), through contract ASI/INAF I/023/05/0.
References
Baldi, A., Molendi, S., Comastri, A., et al. 2002, ApJ, 564,
190B
Cutri, R.M., Skrutskie, M.F., Van Dyk, S. et al. 2003,
University of Massachusetts and Infrared Processing
and Analysis Center,
Institute of
Technology)
(IPAC/California
Table 4. The XMM-Newton Catalogue of sources in the region of SAX J1748.2-2808. The statistical error in the source positions varies in the range 0.5 -- 0.8′′.
Count rates (in units of ks−1) are MOS1+MOS2+pn rates (see text) observed in the following energy ranges: S = 0.5 -- 2 keV, M = 2 -- 5 keV, H = 5 -- 10 keV. The
hardness ratios are defined as follows: HR1 = (M-S)/(M+S) and HR2 = (H-M)/(H+M). For the brightest sources (the first seven), the observed fluxes reported
here, in units of 10−13 erg cm−2 s−1, have been directly obtained from the spectral fits (see Table 2). For the fainter sources, observed fluxes (2 -- 10 keV) have
been calculated as reported in the text, applying a conversion factor to the combined EPIC rate (M+H). For sources detected only in the soft band (S) or
only in the hard band (H), the fluxes (not corrected for absorption) in the 0.5 -- 2 keV or in the 5 -- 10 keV band have been derived and marked with "S" or "H"
respectively. A search within 4" from the source positions in the SIMBAD data base resulted in the positional matches listed in the last column (Notes)
Source R.A.
ID
1
2
3
4
5
6
7
8
9
10
11
12
(J2000)
267.200047
266.817367
267.070445
266.845425
267.242974
266.878546
267.149164
267.017602
267.058541
267.245308
267.202258
267.081919
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
267.071020
267.118308
267.212103
267.162129
266.943950
266.884630
267.215469
267.170685
267.144386
267.021045
266.951207
267.129677
266.982671
267.103188
267.169872
267.044643
267.056520
267.057854
267.303578
Dec.
(J2000)
-28.211022
-28.180065
-28.130656
-28.151610
-28.239226
-28.229876
-27.938457
-28.245968
-28.272493
-28.188918
-28.016414
-28.124028
-28.262540
-28.158466
-28.245925
-28.176110
-28.150689
-28.148614
-28.180396
-28.026434
-27.926348
-28.117880
-28.065710
-28.197335
-28.052198
-28.222073
-28.306960
-28.308560
-28.045786
-28.296586
-28.077256
S
M
H
HR1
HR2
Flux
Notesa
15.14±0.67
2.38±0.49
−
−
7.55±0.59
16.66±0.99
5.16±0.53
0.47±0.23
−
4.45±0.42
7.89±0.58
−
3.17±0.40
1.45±0.22
5.32±0.49
0.31±0.15
−
−
1.27±0.23
−
1.64±0.35
−
−
−
1.85±0.27
−
1.18±0.32
3.01±0.40
−
1.61±0.31
1.42±0.26
23.76±0.92
66.02±2.27
5.60±0.38
20.60±1.28
6.82±0.67
0.84±0.47
5.53±0.66
3.11±0.45
3.37±0.51
0.96±0.30
−
5.99±0.47
29.30±1.44
10.87±0.46
19.03±1.14
2.38±0.36
−
1.31±0.30
5.98±0.49
5.39±0.52
−
−
0.49±0.21
3.15±0.31
1.70±0.43
0.85±0.22
−
0.52±0.23
−
−
−
0.58±0.29
0.98±0.48
−
−
−
−
−
−
−
−
−
−
−
−
−
0.78±0.18
1.33±0.26
2.17±0.39
−
0.69±0.21
−
1.69±0.22
1.83±0.27
0.87±0.19
−
0.87±0.29
−
−
0.86±0.17
−
−
0.22±0.03
0.93±0.01
1.0
1.0
-0.05±0.06
-0.90±0.05
0.03±0.08
0.74±0.12
1.0
-0.65±0.10
-1.0
1.0
-0.30±0.13
-0.26±0.14
-1.0
0.25±0.31
−
−
-1.0
1.0
-0.25±0.25
−
−
−
-1.0
−
-1.0
-1.0
−
-1.0
-1.0
-0.60±0.03
-0.39±0.03
0.32±0.04
-0.04±0.04
-0.48±0.07
-1.0
-0.62±0.08
0.32±0.07
0.23±0.08
-0.62±0.26
−
0.73±0.10
-1.0
-1.0
−
0.20±0.24
1.0
1.0
−
0.091±0.29
-0.56±0.40
1.0
1.0
1.0
−
1.0
−
−
1.0
−
−
5
20
6.8
30
2
0.4
1.4
4
4
0.5
0.2 (S)
2 (H)
Transient (1)
"main" source SAX J1748.2-2808
G0.9+0.1 PWN
SBM80 (2); variable
HD 316290
faint source near SAX J1748.2-2808;
[RHI84] 10-672;sp.type M6 (3)
HD 161824
0.8
0.4
0.2 (S)
0.5
0.8 (H)
1.4 (H)
0.04 (S)
0.5
0.4
1.1 (H)
1.2 (H)
0.5 (H)
0.06 (S)
0.5 (H)
0.04 (S)
0.09 (S) TYC2 929 (4)
0.5 (H)
0.05 (S)
0.05 (S)
aNumbers in parentheses are the following references: (1) Sidoli & Mereghetti 2003; (2) Sidoli, Mereghetti & Belloni 2001b; (3) Raharto et al., 1984; (4) Hog et al. 2000
8
L
.
S
i
d
o
l
i
e
t
a
l
.
:
U
n
v
e
i
l
i
n
g
t
h
e
n
a
t
u
r
e
o
f
S
A
X
J
1
7
4
8
.
2
-
2
8
0
8
w
i
t
h
X
M
M
N
e
w
t
o
n
-
|
astro-ph/0503204 | 1 | 0503 | 2005-03-09T00:16:24 | The Origin of the Spatial Distribution of X-ray luminous AGN in Massive Galaxy Clusters | [
"astro-ph"
] | We study the spatial distribution of a 95% complete sample of 508 X-ray point sources (XPS) detected in the 0.5-2.0 keV band in Chandra ACIS-I observations of 51 massive galaxy clusters found in the MACS survey. Covering the redshift range z=0.3-0.7, our cluster sample is statistically complete and comprises all MACS clusters with X-ray luminosities in excess of 4.5 x 10^44 erg/s (0.1-2.4 keV, h_0=0.7, LCDM). Also studied are 20 control fields that do not contain clusters. We find the XPS surface density, computed in the cluster restframe, to exhibit a pronounced excess within 3.5 Mpc of the cluster centers. The excess, believed to be caused by AGN in the cluster, is significant at the 8.0 sigma confidence level compared to the XPS density observed at the field edges. No significant central excess is found in the control fields. To investigate the physical origin of the AGN excess, we study the radial AGN density profile for a subset of 24 virialized clusters. We find a pronounced central spike (r<0.5 Mpc), followed by a depletion region at about 1.5 Mpc, and a broad secondary excess centered at approximately the virial radius of the host clusters (~2.5 Mpc). We present evidence that the central AGN excess reflects increased nuclear activity triggered by close encounters between infalling galaxies and the giant cD-type elliptical occupying the very cluster center. By contrast, the secondary excess at the cluster-field interface is likely due to black holes being fueled by galaxy mergers. In-depth spectroscopic and photometric follow-up observations of the optical counterparts of the XPS in a subset of our sample are being conducted to confirm this picture. | astro-ph | astro-ph | APJL, ACCEPTED
Preprint typeset using LATEX style emulateapj v. 6/22/04
5
0
0
2
r
a
M
9
1
v
4
0
2
3
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
THE ORIGIN OF THE SPATIAL DISTRIBUTION OF X-RAY LUMINOUS AGN IN MASSIVE GALAXY CLUSTERS
JOSHUA T. RUDERMAN1 AND HARALD EBELING
Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive, Honolulu, HI 96822
ApJL, accepted
ABSTRACT
We study the spatial distribution of a 95% complete sample of 508 X-ray point sources (XPS) detected
in the 0.5 -- 2.0 keV band in Chandra ACIS-I observations of 51 massive galaxy clusters found in the MACS
survey. Covering the redshift range z = 0.3 − 0.7, our cluster sample is statistically complete and comprises
all MACS clusters with X-ray luminosities in excess of 4.5 × 1044 erg s−1 (0.1 -- 2.4 keV, h0 = 0.7, ΛCDM).
Also studied are 20 control fields that do not contain clusters. We find the XPS surface density, computed
in the cluster restframe, to exhibit a pronounced excess within 3.5 Mpc of the cluster centers. The excess,
believed to be caused by AGN in the cluster, is significant at the 8.0σ confidence level compared to the XPS
density observed at the field edges. No significant central excess is found in the control fields. To investigate
the physical origin of the AGN excess, we study the radial AGN density profile for a subset of 24 virialized
clusters. We find a pronounced central spike (r < 0.5 Mpc), followed by a depletion region at about 1.5 Mpc,
and a broad secondary excess centered at approximately the virial radius of the host clusters (≈ 2.5 Mpc). We
present evidence that the central AGN excess reflects increased nuclear activity triggered by close encounters
between infalling galaxies and the giant cD-type elliptical occupying the very cluster center. By contrast, the
secondary excess at the cluster-field interface is likely due to black holes being fueled by galaxy mergers. In-
depth spectroscopic and photometric follow-up observations of the optical counterparts of the XPS in a subset
of our sample are being conducted to confirm this picture.
Subject headings: galaxies: active -- galaxies: clusters: general -- galaxies: evolution -- galaxies: interactions
-- X-rays: galaxies -- X-rays: galaxies: clusters
1. INTRODUCTION
The abundance and properties of AGN in galaxy clusters
are important diagnostics for studies of cluster formation and
galaxy evolution, but are difficult to measure as a large num-
ber of homeogenously selected systems is required for a sta-
tistically robust result. Massive galaxy clusters should be pre-
ferred targets for this kind of research as they constitute the
largest reservoirs of galaxies. On the other hand, the galaxy
population specifically of evolved clusters is overwhelmingly
dominated by elliptical galaxies which, at least in the local
universe, are gas-poor and thus do not typically exhibit nu-
clear activity. The situation is complicated further by the fact
that, until recently, observational studies of the AGN distribu-
tion in clusters and the field have based AGN identifications
almost exclusively on characteristic emission lines in the op-
tical part of the galaxy spectrum, making them insensitive to
a hypothetical population of extremely obscured AGN.
In the past few years, several studies have taken a com-
plementary approach by using X-ray observations, specifi-
cally the surface density of X-ray point sources (XPS) around
galaxy clusters,
to quantify the AGN content of clusters
relative to that of the field. Evidence of an XPS excess
has been presented for several clusters over a wide range
of redshifts and X-ray luminosities (Henry & Briel 1991;
Cappi et al. 2001; Molnar et al. 2002; Sun & Murray 2002;
Cappelluti et al. 2004). If the point sources detected in these
fields are indeed at the cluster redshift, their X-ray luminosi-
ties suggest that the excess is caused by AGN in the clus-
ter and can thus be used to map the cluster's AGN distribu-
tion. Optical follow-up observations performed in the field of
A2104 at z = 0.154 (Martini et al. 2002) confirm this picture
1 also Departments of Physics and Mathematics, Stanford University,
Stanford, CA 94305
and reveal a possibly considerable fraction of these galaxies
to lack the spectroscopic characteristics of AGN, suggesting
that clusters may contain a large subset of optically obscured
AGN.
Using Chandra ACIS-I data we here present the results
of the first systematic study of the XPS content of a large,
representative sample of massive galaxy clusters, to test the
findings of earlier studies obtained for individual clusters.
The large number of XPS detections obtained in this work
allows us to construct the first resolved radial profile of XPS
in the cluster rest frame, and to characterize the spatial XPS
distribution as well as its likely physical origin. Throughout
we assume a ΛCDM cosmology with h0=0.7, Λ=0.7, and
Ω=0.3.
2. FIELD SELECTION
We analyze Chandra ACIS data (front-illuminated chips
only) for all 51 MACS clusters (MACS = MAssive Cluster
Survey) (Ebeling, Edge & Henry 2001) at z = 0.3 − 0.7 that
have been observed before Jan 15, 2005 and feature X-ray
luminosities in excess of 4.5 × 1044 erg s−1 (0.1 -- 2.4 keV).
The total geometric area covered by these observations is
5.73 deg2. An identical analysis is performed for ACIS-I
data of 20 control fields, i.e. observations that did not target
galaxy clusters, covering a total of 1.98 deg2. On-axis
exposure times range from 10 ks to 87 ks. We reduce the raw
data using CIAO software version 3.02, removing systematic
instrumental effects and background flares. Merged datasets
are created for the nine clusters observed twice and three
clusters observed three times (see Fig. 1 for an example). To
account for exposure-time variations across the ACIS field
of view (vignetting, chip gaps, etc) we generate effective
exposure maps for the observed XPS peak energy of 1.17 keV.
2
Ruderman and Ebeling
30
30
25
25
20
20
15
15
10
10
05
05
38° 00′
38° 00′
37° 55′
37° 55′
)
)
0
0
0
0
0
0
2
2
(
(
n
n
o
o
i
i
t
t
a
a
n
n
i
i
l
l
c
c
e
e
D
D
23m 00s
23m 00s
30.07s
30.07s
22m 00s
22m 00s
Right Ascension (2000)
Right Ascension (2000)
30.05s
30.05s
21m 00s
21m 00s
16h 0.09s
16h 0.09s
FIG. 1. -- Merged dataset constructed from the three ACIS observations of
the relaxed cluster MACS J1621.3+3810 at z = 0.46 (Edge et al. 2003). Red
ellipses mark the point sources detected with Celldetect using a 3σ detection
threshold. The green rectangles show the orientation of the front-illuminated
chips for the three individual exposures.
3. DATA REDUCTION
In order to maximize the detection efficiency we run the
Celldetect algorithm in the 0.5 -- 2.0 keV range2; only sources
with detection significance of at least 3σ above the local X-ray
background are kept. For each cluster we use the appropriate
ACIS exposure map to identify and remove spurious detec-
tions at the chip edges, leaving 910 point source detections in
the cluster fields and 509 detections in the control fields.
Source count rates are converted to unabsorbed energy
fluxes in the 0.5 -- 2.0 keV energy band assuming the Galactic
nH value and a power law with a spectral index of Γ = 1.7,
consistent with the observed stacked XPS spectrum3. Spa-
tial variations of the detector characteristics are taken into ac-
count in the conversion by computing Redistribution Matrix
Files (RMF) and Auxiliary Response Files (ARF) for each
source individually. Using a power-law model of the XPS
log N − log S obtained from a maximum-likelihood fit to the
surface-density distribution of the XPS detected in 23 clus-
ter observation of duration 17.5 -- 20.5 ks, the most common
exposure time, we compute instrumental flux limits of 95%
completeness for each cluster scaled to individual exposure
times. We find the source lists for all fields to be complete to
a global flux limit of 1.25 × 10−14 erg s−1 cm−2, correspond-
ing to an on-axis exposure time of 10 ks. For point sources
in cluster fields, we convert unabsorbed X-ray fluxes to lumi-
nosities assuming the point sources to be at the redshifts of the
respective target cluster. Details of the properties of the target
clusters as well as of the XPS population found in the cor-
responding ACIS-I fields will be presented in a forthcoming
2 This bandpass is selected to maximize the signal-to-noise ratio for
sources with AGN-like X-ray spectra
3 The details of the chosen spectral model are not critical as, in this energy
band, the observed flux depends only weakly on the chosen power law slope
(Cappi et al. 2001).
0.8
0.8
0.6
0.6
)
)
2
2
−
−
c
c
p
p
M
M
(
(
n
n
0.4
0.4
0.2
0.2
0.0
0.0
0
0
0.08
0.06
0.04
0.02
0.00
0
)
2
−
n
i
m
c
r
a
(
n
1
1
2
2
3
3
r (Mpc)
4
4
5
5
6
6
7
7
5
10
r (arcmin)
15
20
FIG. 2. -- Radial profiles of the XPS surface density in 51 cluster fields
(top) and 20 control fields (bottom). In the cluster fields, the XPS density
is constant beyond distances of about 3.5 Mpc from the cluster center. The
red bar marks this background level and its 1σ error as derived from a fit to
the data in the 4 -- 7 Mpc region. Within 3.5 Mpc a highly significant (8.0σ)
excess is observed in the cluster fields. By contrast the XPS surface density
in the control fields exhibits no excess. Both graphs probe similar angular
scales because, at the redshifts of our clusters, one Mpc corresponds to 2.3 to
3.7 arcminutes.
paper (Ebeling et al., in preparation).
4. DISTRIBUTION OF POINT SOURCES
We compute the metric distance from the cluster center for
all XPS detected in cluster fields, assuming again that the XPS
are at the cluster redshift. A radial profile of their surface den-
sity is then constructed by binning the radial distances for the
XPS from all 51 clusters, eliminating XPS fainter than the flux
limit of the respective observation (this leaves 508 XPS), and
dividing the XPS counts in each bin by the appropriate area.
The result is shown in Fig. 2 (top). The graph at the bottom of
the same figure shows a similar histogram for the 256 XPS de-
tected above the individual flux limits of the 20 control fields,
using the angular distances from the observation target posi-
tion which corresponds to the location on the detector where
cluster centers fall in cluster fields. For both panels of Fig. 2,
as well as for Fig. 3, the value of the surface density in each
bin is computed as n = Pi Ni/ Pi Ai, with Ni and Ai be-
ing the XPS number and area of the respective annulus for the
ith cluster; the shown 1σ error bars assume Poisson statistics
using the analytic approximations of Ebeling (2003).
X-ray luminous AGN in massive galaxy clusters
3
The significance of the central XPS excess observed
in the cluster fields can be computed from the difference
between the XPS surface density observed in the inner 3.5
Mpc of the cluster fields and the one observed in the outer
4 -- 7 Mpc annulus, which we assume to be the background
XPS density. We find the excess to be significant at the
8.0σ confidence level. No central excess is observed in the
radial XPS density profile for the control fields, ruling out
a systematic instrumental effect or processing error. Note
that the background per Mpc2 implied by the control fields
is higher than the one measured in the outer regions of
the cluster fields. We believe that this discrepancy reflects
differences in the large-scale environment between the cluster
and control samples which we will discuss in a future paper
(Ebeling et al., in preparation).
5. CORRELATION WITH CLUSTER MORPHOLOGY
In order to gain insight into the physical origin of the ra-
dial XPS distribution shown in Fig. 2 (top) we crudely clas-
sify our 51 MACS clusters as relaxed or disturbed based
on their X-ray morphology. Using Chandra images adap-
tively smoothed with the ASMOOTH algorithm (Ebeling
White & Rangarajan2005), we consider a cluster to be relaxed
only if its X-ray surface-brightness distribution exhibits the
following characteristics: a) on small scales, a pronounced
central peak typical of a cooling core, and b) on large scales,
near-perfect spherical symmetry. The combination of these
criteria ensures that the clusters thus selected have not under-
gone a major merger event in several Gyr and can be consid-
ered virialized. For these systems the metric distance used to
construct the histogram shown in Fig. 2 (top) corresponds to a
well defined cluster-centric radius which probes values of the
basic cluster properties -- such as gas temperature and den-
sity, or galaxy volocity dispersion -- that show little variation
with azimuthal angle.
Our morphological classification splits our XPS list into
two X-ray flux limited subsamples, each containing 254 XPS,
detected in the fields of 24 relaxed or 27 disturbed systems,
respectively. Fig. 3 shows the radial XPS density profiles
for both subsets separately. Although both profiles exhibit a
prominent excess of XPS within approximately 3.5 Mpc of
the cluster center -- significant at 6.3 and 4.7σ confidence
for relaxed and disturbed systems, respectively -- the two
distributions are markedly different (at 3.0σ signficance) with
only the spherically symmetric, relaxed clusters showing the
double peak apparent already in Fig. 2 (top).
6. CORRELATION WITH CLUSTER REDSHIFT
We attempt to also test how the excess depends on cluster
redshift. In doing so, it is critical that the luminosities of the
AGN probed at different redshifts are comparable, i.e., com-
pleteness to a universal luminosity limit needs to be ensured.
For a first, crude check of any redshift dependence we divide
our cluster sample into two subsets, comprising 29 clusters at
z = 0.3 − 0.45 and 21 clusters at z = 0.45 − 0.6. Converting
the flux limits of the different observations into luminosities,
we find that our XPS lists for the z = 0.3 − 0.6 redshift range
are 95% complete to a global luminosity limit of 1.1 × 1043
erg s−1 (0.5 -- 2.0 keV). Applying this luminosity cut reduces
our XPS sample to 87 and 119 sources in the z = 0.3 − 0.45
and z = 0.45 − 0.6 cluster fields, respectively. Although the
data are suggestive of an increase of the excess with redshift,
)
)
2
2
−
−
c
c
p
p
M
M
(
(
n
n
)
)
2
2
−
−
c
c
p
p
M
M
(
(
n
n
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
0.2
0.2
0.0
0.0
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
0.2
0.2
0.0
0.0
0
0
1
1
2
2
3
3
r (Mpc)
r (Mpc)
4
4
5
5
6
6
7
7
FIG. 3. -- Radial profiles of the XPS surface density in the fields of 24
relaxed clusters (top) and in the fields of 27 disturbed clusters (bottom). The
profile for relaxed clusters features a pronounced central peak at r < 0.5
Mpc, a depletion region at 0.5 -- 1.5 Mpc, and a broad secondary excess at
approximately the virial radius of 2 -- 3 Mpc. By contrast, the XPS density
profile for disturbed clusters shows a smoothly distributed excess.
the difference between the two radial profiles is not significant
(1.5σ) at the current sizes of our subsamples.
7. DISCUSSION
7.1. XPS nature
The X-ray luminosities referred to in the previous section
can be used to characterize the nature of the point sources
causing the observed excess. As a result of the, in general,
moderate exposure times of the ACIS observations used in
our study, we find all of the XPS in our flux-complete sample
to feature X-ray luminosities well in excess of 3 × 1042 erg
s−1 (0.5 -- 8.0 keV). Since this value represents a tight upper
limit to the X-ray luminosity of starburst galaxies (Bauer et al.
2004) we conclude that the observed XPS excess can be at-
tributed to AGN at the cluster redshift.
The alternative explanation that the excess is caused by
gravitational lensing of background AGN (which would then
have to be yet more X-ray luminous) can be ruled out because
of the large area over which the excess is observed. The am-
plification provided by lensing can account at best for a small
fraction of the observed excess at r < 0.5 Mpc, and is en-
tirely negligible at the angular distances corresponding to the
radius at which the secondary excess is observed (Smith et al.
2002). In addition, studies searching for an XPS excess in
cluster fields observed to much lower X-ray fluxes indicate
that the faint end of the log N − log S distribution of the ex-
cess is inconsistent with the one predicted for gravitationally
4
Ruderman and Ebeling
lensed sources (Cappi et al. 2001). We therefore conclude that
the excess is due to AGN in the cluster.
7.2. XPS distribution
We now discuss the possible origin of the spatial XPS dis-
tribution, which is characterized by a central excess within 0.5
Mpc of the cluster core, a depletion region around 1.5 Mpc,
and a broad secondary excess observed at about the virial ra-
dius of the relaxed clusters. Although present in the fields of
all clusters, the excess at r < 0.5 Mpc is much more pro-
nounced for the relaxed systems, all of which possess central
cooling cores dominated by massive cD galaxies. Closer in-
spection of the XPS distribution reveals that the difference
in this first radial bin between relaxed and disturbed systems
originates at yet smaller cluster-centric distances. We detect
six XPS within 250 kpc of the cD galaxies in virialized clus-
ters and none within the same distance of the cores of dis-
turbed systems. Two of these six XPS in the very heart of
virialized clusters can be unambiguously identified as AGN
in the cD itself; for the other four cases we propose that the
observed X-ray emission is due to nuclear activity triggered
by close encounters between infalling cluster galaxies and the
cD. Extending this picture to the remainder of the r < 0.5
Mpc bin, we predict that the central excess is due to galaxy
mergers and tidal interactions involving the giant elliptical
galaxies dominating the centers of MACS clusters.
The AGN depletion zone in relaxed clusters at radii around
1.5 Mpc and the following broad excess at roughly the virial
radius can be explained by a slightly different mechanism.
The rise of the observed XPS surface density at 2 − 3 Mpc
suggests increased AGN activity at the cluster-field interface,
which is easily explained by merger-induced accretion onto
massive black holes, favoured by the increase in galaxy den-
sity compared to the field and the presence of relatively gas-
rich galaxies in this transition region. Since successful merg-
ers require low collision velocities of typically less than 300
km s−1 (Binney & Tremaine 1987) such mergers become ex-
tremely rare closer to the cluster core where, for MACS clus-
ters, the galaxy velocity dispersion reaches and exceeds 1000
km s−1 (Barrett et al., in preparation). As a result, the AGN
density is dramatically reduced at intermediate radii (0.5 --
2 Mpc).
In addition, activity triggered at larger radii will
cease as infalling galaxies approach the cluster core since the
timescale for the depletion of the accretion disk is substan-
tially shorter than the cluster crossing time. The combination
of these two effects ensures that the excess is confined to the
infall region where AGN activity is triggered.
We expect the same physical mechanisms to cause AGN ac-
tivity in disturbed clusters. However, since disturbed clusters
lack well defined cores as well as spherical symmetry, the ex-
cess is spread much more evenly over the 0 − 3.5 Mpc range,
as observed (Fig. 3, bottom).
8. CONCLUSIONS
We present results of the first study of the X-ray point
source content of a statistically well defined sample of galaxy
clusters, using Chandra/ACIS observations of 51 MACS clus-
ters at z = 0.3 − 0.7. We detect an overall 8.0σ significant ex-
cess in the point source density within 3.5 Mpc of cluster cen-
ters, which we argue is caused by AGN in the cluster targets
of these observations. We also present the first resolved radial
profile of the excess and find it to depend significantly on clus-
ter morphology. Making use of the simple geometry afforded
by relaxed clusters we are able to identify two distinct com-
ponents to the excess, namely a central excess of AGN which
we believe to be due to galaxy interactions involving the giant
ellipticals near the cluster core, and a broad secondary excess
at about the virial radius. We propose that the AGN activity
causing the secondary excess is triggered by galaxy mergers,
which are most likely to occur in the low-energy collisions
favoured in the cluster-field transition region. The lack of
AGN at intermediate radii can be explained by the low merger
probability at high relative velocities and the shortness of the
accretion timescale compared to the cluster crossing time.
Our results confirm those of past studies which have de-
tected point source excesses in cluster fields and provide a
novel way to probe the dynamics of AGN production by
galaxy interactions in massive clusters. Future work will fo-
cus on the optical counterparts of X-ray point sources in se-
lected MACS clusters to unambiguously identify redshifts and
X-ray emission processes, thus testing conclusively our hy-
pothesis that the excess is due to AGN cluster members.
The authors gratefully acknowledge financial support from
NASA grant NAG 5-8253 and SAO grant GO2-3168X (HE),
and from the National Science Foundation's Research Expe-
riences for Undergraduates program (JTR). We thank every-
body at the Chandra X-Ray Center for their contribution to
a spectacularly successful satellite mission and specifically
for maintaining the CIAO software package. Thanks also
to Joshua Barnes for a very informative discussion about the
physics and consequences of galaxy mergers, and to the ref-
eree, Stefano Ettori, for helpful comments and criticism.
REFERENCES
Bauer, F.E., Alexander, D.M., Brandt, W.N., Schneider, D.P., Treister, E.,
Edge, A.C., Ebeling, H., Bremer, M., Rottgering, H., van Haarlem, M.P.,
Rengelink, R., Courtney, N.J.D. 2003, MNRAS, 339, 913
Henry, J.P., & Briel, U.G. 1991, A&A, 246, 14L
Martini, P., Kelson, D.D., Mulchaey, J.S. & Trager, S.C. 2002, ApJ, 576, L109
Molnar, M.S., Hughes, J.P., Donahue, M. & Joy, M. 2002, ApJ, 573, L91
Smith, G.P. et al. 2002, MNRAS, 330, 1
Sun, M., & Murray, S.S. 2002, ApJ, 577, 139
Hornschemeier, A.E., & Garmire, G.P. 2004, ApJ, 128, 2048
Binney, J. & Tremaine S. 1987, Galactic Dynamics, Princeton: Princeton
University Press
Cappelluti, N., Cappi, M., Dadina, M., Malaguti, G., Branchesi, M., DElia,
V., & Palumbo, G.G.C. 2004, A&A, in press
Cappi, M. et al. 2001, ApJ, 548, 624
Ebeling, H., Edge, A.C., & Henry, J.P. 2001, ApJ, 553, 668
Ebeling, H. 2003, MNRAS, 340, 1269
Ebeling, H., White, D.A., & Rangarajan V.N. 2005, MNRAS, submitted
|
astro-ph/0007087 | 2 | 0007 | 2001-04-03T17:50:37 | The transition to nonlinearity and new constraints on biasing | [
"astro-ph"
] | We present two new dynamical tests of the biasing hypothesis. The first is based on the amplitude and the shape of the galaxy-galaxy correlation function, $\xi_g(r)$, where $r$ is the separation of the galaxy pair. The second test uses the mean relative peculiar velocity for galaxy pairs, $\vs(r)$. This quantity is a measure of the rate of growth of clustering and it is related to the two-point correlation function for the matter density fluctuations, $\xi(r)$. Under the assumption that galaxies trace the mass ($\xi_g = \xi$), the expected relative velocity can be calculated directly from the observed galaxy clustering. The above assumption can be tested by confronting the expected $\vs$ with direct measurements from velocity-distance surveys. Both our methods are checked against N-body experiments and then compared with the $\xi_g(r)$ and $\vs$ estimated from the {\sc APM} galaxy survey and the Mark III catalogue, respectively. Our results suggest that cosmological density parameter is low, $\Omega_m \approx 0.3$, and that the {\sc APM} galaxies trace the mass at separations $r \ga 5 \Mlu$, where $h$ is the Hubble constant in units of 100 km s$^{-1}$Mpc. The present results agree with earlier studies, based on comparing higher order correlations in the {\sc APM} with weakly non-linear perturbation theory. Both approaches constrain the linear bias factor to be within 20% of unity. If the existence of the feature we identified in the {\sc APM} $\xi_g(r)$ -- the inflection point near $\xi_g = 1$ -- is confirmed by more accurate surveys, we may have discovered gravity's smoking gun: the long awaited ``shoulder'' in $\xi$, generated by gravitational dynamics and predicted by Gott and Rees 25 years ago. | astro-ph | astro-ph |
The transition to nonlinearity and new constraints on
biasing
Roman Juszkiewicz1,2,3 & Enrique Gaztanaga4,5
1 J. Kepler Astronomical Center, 65-265 Zielona G´ora, Poland
2 D´epartement de Physique Th´eorique, Universit´e de Gen`eve, CH-1211
Gen`eve, Switzerland
3 N. Copernicus Astronomical Center, 00-716 Warsaw, Poland
4 INAOE, Astrofisica, Tonantzintla, Apdo Postal 216 y 51, Puebla 7200,
Mexico
5 Institut d'Estudis Espacials de Catalunya, IEEC/CSIC, Edf.
Nexus-201 - c/ Gran Capitan 2-4, 08034 Barcelona, Spain
Abstract.
We present two new dynamical tests of the biasing hypothesis. The
first is based on the amplitude and the shape of the galaxy-galaxy cor-
relation function, ξg(r), where r is the separation of the galaxy pair.
The second test uses the mean relative peculiar velocity for galaxy pairs,
v12(r). This quantity is a measure of the rate of growth of clustering
and it is related to the two-point correlation function for the matter den-
sity fluctuations, ξ(r). Under the assumption that galaxies trace the mass
(ξg = ξ), the expected relative velocity can be calculated directly from the
observed galaxy clustering. The above assumption can be tested by con-
fronting the expected v12 with direct measurements from velocity-distance
surveys. Both our methods are checked against N-body experiments and
then compared with the ξg(r) and v12 estimated from the APM galaxy
survey and the Mark III catalogue, respectively. Our results suggest that
cosmological density parameter is low, Ωm ≈ 0.3, and that the APM
galaxies trace the mass at separations r >∼ 5 h−1Mpc, where h is the
Hubble constant in units of 100 km s−1Mpc. The present results agree
with earlier studies, based on comparing higher order correlations in the
APM with weakly non-linear perturbation theory. Both approaches con-
strain the linear bias factor to be within 20% of unity. If the existence
of the feature we identified in the APM ξg(r) -- the inflection point near
ξg = 1 -- is confirmed by more accurate surveys, we may have discovered
gravity's smoking gun: the long awaited "shoulder" in ξ, generated by
gravitational dynamics and predicted by Gott and Rees 25 years ago.
1.
Introduction
We have been recently working on two projects, directly related to the topic
of this Meeting -- the transition to nonlinearity in gravitational clustering. We
1
have therefore decided to present some of our results here. The first idea -- to
use the inflection point in the galaxy-galaxy correlation function to constrain
biasing is in a more mature state than the second, which uses measurements of
the relative motions in pairs of galaxies. Accordingly, the first project is almost
a paper -- it has been submitted for publication in Monthly Notices. The results
of our work on the second project are still far from completion but ripe enough
to be presented and discussed.
1.1. The CDM crisis and biasing
The concept that galaxies may not be fair tracers of the mass distribution was in-
troduced in the early eighties, in part in response to the observation that galaxies
of different morphological types have different spatial distributions, hence they
cannot all trace the mass (there are two excellent reviews on the subject: Strauss
& Willick 1995 and Hamilton 1998). However, there was also another reason:
to "satisfy the theoretical desire for a flat universe" (Davis et al. 1985, p.391).
More precisely, biasing was introduced to reconcile the observations with the
predictions of the Einstein-de Sitter cold dark matter (CDM) dominated model.
At the time, it seemed that just a simple rescaling of the overall clustering ampli-
tude by setting ξg(r) = b2ξ(r), where b ≈ 2 might do the job (Davis et al. 1985).
However, very soon thereafter, it became clear that this is not enough: while the
unbiased (b = 1) ξ(r) had too large an amplitude at small r, the biased model
did not have enough large-scale power to explain the observed bulk motions (Vit-
torio et al. 1987). A similar conclusion could be drawn form comparison of the
relative amplitude of clustering on large and small scales (eg Maddox et al 1990).
The problem with the shape of ξ(r) became explicit when measurements of ξg(r)
showed that the optically selected galaxies follow an almost perfect power law
over nearly three orders of magnitude in separation. This result disagrees with
N-body simulations. The standard (Ωm = 1) CDM model as well as its various
modifications, including Ωm < 1 and a possible non-zero cosmological constant,
fail to match the observed power law (see Fig 11-12 in Gaztanaga 1995, Jenk-
ins et al. 1998; most of these problems were already diagnosed by Davis et al.
1985). Two alternative ways out of this impasse were recently discussed by Rees
(1999) and Peebles (1999). We believe that it will be helpful to discuss both
approaches because their existence provides the motivation for our work.
1.2. Environmental cosmology
A possible response to the CDM crisis is to build a model where simple phe-
nomena, like the power-law behavior of ξg are much more complicated than they
seem. In particular, one can explore the possibility that the emergence of large
scale structure is not driven by gravity alone but by "environmental cosmology"
-- a complex mixture of gravity, star formation and dissipative hydrodynamics
(Rees 1999). A phenomenological formalism, appropriate for this approach was
recently proposed Dekel & Lahav (1999). According to the old, "linear biasing"
prescription, at each smoothing scale, the galaxy- and the dark matter-density
fields, δg and δ, are related by the linear function
δg = b δ ,
2
(1)
where b is a time- and scale-independent constant. In the new picture, the rela-
tionship between the two fields is neither linear nor deterministic. The biasing
factor, here defined as the ratio of the rms values of the two fields, b = σg/σ is
allowed to depend on the smoothing scale as well as on the cosmological time.
A convenient measure of the stochasticity in the relationship between the two
fields is the cross-correlation coefficient,
R ≡
hδδgi
σgσ
;
R ≤ 1 .
(2)
The special case R = 1 describes the deterministic bias, while R = 0 corresponds
to the "maximum stochasticity", or the case when the two fields are completely
uncorrelated. The parameters R and b describe he biasing and stochasticity in
the galaxy distribution relative to the spatial mass distribution. When referring
bIRAS for
to specific galaxy surveys, we will sometimes use subscripts, e.g.
the IRAS survey. These quantities should be distinguished from the bias and
stochasticity measures for two classes of galaxies of different morphological types,
e.g.
for early (subscript e) and late (subscript l) galaxies: bel ≡ σe/σl, and
Rel ≡ hδeδli(σeσl)−1.
The above parameters can be estimated or constrained by more or less di-
rect measurements from galaxy redshift surveys, peculiar velocity data and other
observations. They can be also studied in semi-analytic theoretical models (Sel-
jak 2000, Scoccimarro et al. 2000), hydrodynamic simulations or semi-analytic
models combined with N-body simulations (eg. Blanton et al. 2000, Somerville
et al. 2001 and references therein).
1.3. Constraints on biasing
If biasing is indeed important and complicated, we should expect that b, bel, R
and Rel are all significantly different from unity and scale-dependent. As we
show below, at sufficiently large scales, r >∼ 10h−1Mpc, there is actually evidence
to the contrary: the admissible deviations of b and R are small and always
comparable to the accuracy of the measurements.
Skewness and higher moments. The strongest constraints on large
(weakly non-linear) scales come from the measurements of the two-, three-
and four-point connected moments of the density field in the APM catalogue.
These constraints are obtained as follows. One assumes that galaxies trace
the mass and the large-scale structure we observe today grew out of small-
amplitude, Gaussian density fluctuations in an expanding, self-gravitating non-
relativistic gas.
If our assumption is correct, by now nonlinear gravitational
instability would have driven the distribution away from gaussianity, generating
skewness and higher connected moments, which can be calculated analytically
(Juszkiewicz et al. 1993, Bernardeau 1994a, 1994b). The assumptions about
initial gaussianity and lack of biasing can then be tested by comparing the an-
alytical predictions with observations. The predictions are in good agreement
with the data from the APM (Gaztanaga 1994, Gaztanaga & Frieman 1994,
Frieman & Gaztanaga 1999) as well as the IRAS PSCz catalogue (Feldman et
al. 2000) and suggest that b(r) is within 20% of unity for r >∼ 10h−1Mpc, or at
linear scales (where the clustering amplitude is less than unity).
3
Redshift distortions. Some of the most radical claims that bel can be as
large as 1.5 to 2 are based on comparisons of the estimates of the strength of
clustering in the 1.2 Jy IRAS catalogue with optical redshift surveys (Strauss &
Willick 1995 and references therein). Indeed, those earlier studies, summarized
by Hamilton (1998), gave a redshift distortion parameter βIRAS = Ω0.6
m /bIRAS =
0.77±0.22, while the average and the standard deviation of the same parameter,
obtained from optical surveys is, according to the same compilation, βoptical =
0.52 ± 0.26,
implying a relative bias boptical/bIRAS = bel ≈ 1.5. The most
recent analysis (Hamilton et al. 2000) of the larger PSCz sample Saunders et al.
2000 gives βIRAS = 0.41 ± 0.13, consistent with bel = 1. Moreover, Hamilton
et al. also conclude that their results are consistent with RIRAS(r) ≈ 1 at
r >∼ 10h−1Mpc. More recently Hamilton and Tegmark (2000) have studied
the large as well as small-scale clustering in real space, reconstructed from the
redshift space data of the PSCz survey and found evidence for scale-dependent
bias; however, the effect is significant at small, strongly-nonlinear scales only.
Large scale flows. Recent measurements of the mean relative pairwise
velocity of galaxies allow an independent estimate of Ωm and the biasing param-
eter. These results are consistent with Ωm ≈ 0.3 and R ≈ b ≈ 1, bel ≈ Rel ≈ 1
at separations r >∼ 5h−1Mpc (Juszkiewicz et al. 2000a).
Weak lensing. The values of b and Ωm, derived from large scale flows
are consistent with recent measurements of cosmic shear correlations based on
the VIRMOS deep imaging survey (Van Waerbeke et al. 2001). On much
smaller scales, r ≈ 1h−1Mpc, weak lensing data from the Sloan Survey give
ΩmR/b ≈ 0.3 (Fischer et al. 1999), again in agreement with the estimates from
the pairwise motions, although unlike the relative velocity approach, the Sloan
results suffer from degeneracies between Ωm and bias.
Qualitative arguments. Qualitatively, strong biasing effects would be
difficult to reconcile with the well known fact that the L∗ galaxies, dwarf galax-
ies and IRAS galaxies have strikingly similar distributions, all avoiding the voids
(Peebles 1993, pp. 638-642). Another empirical argument against biasing is pro-
vided by the universal character of the observed Tully-Fisher and fundamental
plane relations (see, for example Binney 1999; Peebles 1999).
Simulations. All simulations of the galaxy formation process, either of
hydrodynamic or semi-analytic variety predict morphological segregation as well
as b and R parameters dependent on scale and time. Since these models are at
a relatively early stage of their development, the details differ from model to
model on small scales and at high redshifts, i.e. exactly where their b and R
parameters are significantly different from unity and where clustering is strongly
nonlinear. However, most of these numerical experiments broadly agree that
with decreasing redshift and increasing scales, b and R approach unity (see eg.
Figure 18 in Somerville et al. 2001 and references therein).
In the end, it may turn out that to explain the observed structure, all
we need is just the plain gravitational instability theory, leaving complex non-
gravitational physics on scales below a Megaparsec to cosmogony, directly in-
volving star formation. We discuss this possibility below.
4
1.4. What you get is what you see
An obvious alternative to environmental cosmology was recently discussed by
Peebles (1999), who pointed out that "as Kuhn has taught us, complex inter-
pretations of simple phenomena have been known to be precursors of paradigm
shifts" and perhaps after fifteen years of attempts to salvage the CDM model
with biasing, it is time to abandon this approach, as well as the biasing hypoth-
esis itself as "another phlogiston" (all quotations in inverted commas are from
Peebles 1999). Instead, one can explore a simpler option, that galaxies trace the
mass distribution, or
ξg = ξ and R = b = 1 ,
(3)
at least for local (low redshift), optically selected galaxies with a broad magni-
tude sampling. This approach rests on the idea that no matter how or where
galaxies form, they must eventually fall into the dominant gravitational wells
and therefore trace the underlying mass distribution (see Peebles 1980, hereafter
LSS; Fry 1996). Our purpose here is to test this idea, using measurements of
relative velocities of pairs of galaxies and the shape of the two-point correlation
function.
1.5. Outline of the paper
In this paper we propose two new tests of the biasing hypothesis, which involve
two measures of clustering. The first is the two-point correlation function of
mass density fluctuations, ξ. The second is a measure of the rate of gravita-
tional clustering -- the relative velocity of particle pairs, v12. We describe our
theoretical model in the next section. Our analytic formulae used to test the
biasing hypothesis are checked against N-body simulations in §3. In §4 we apply
our tests to the APM galaxy survey. Finally, in §5 we summarize and discuss
our results.
2. Theory
2.1. The relative velocity
The relative pairwise velocity v12 was introduced in the context of the BBGKY
theory (Davis and Peebles 1977), describing the dynamical evolution of a col-
lection of particles interacting through gravity. In this discrete picture, ~v12 is
defined as the mean value of the peculiar velocity difference of a particle pair at
separation ~r (LSS, Eq. 71.4). In the fluid limit, its analogue is the pair-density
weighted relative velocity Fisher et al. 1994, Juszkiewicz et al. 1998,
~v12(r) =
h(~v1 − ~v2)(1 + δ1)(1 + δ2)i
1 + ξ(r)
,
(4)
where ~vA and δA = ρA/hρi − 1 are the peculiar velocity and fractional density
contrast of matter at a point A = 1, 2, . . .. The brackets h. . .i denote ensemble
averages for pairs of points at a fixed separation r = ~r1 −~r2, while ξ(r) = hδ1δ2i.
In gravitational instability theory, the magnitude of ~v12(r) is related to ξ(r)
through the pair conservation equation (LSS, Eq. 71.6).
5
Recently Juszkiewicz et al. (1999, hereafter JSD) have proposed an approx-
imate solution of the pair conservation equation. Using Eulerian perturbation
theory, they solved the equation for v12(r) to second order in ξ. They also pro-
posed an interpolation between their large-r perturbative limit, and the well
known small separation limit -- the stable clustering solution, v12(r) = −Hr,
where H is the Hubble constant (LSS, §71). The resulting ansatz is given by
v12(r) = − 2
¯ξ(r) = (3/r3) Z r
0
3 Hf r¯¯ξ(r)[1 + α ¯¯ξ(r)] ,
ξ(x) x2 dx ≡ ¯¯ξ(r) [ 1 + ξ(r) ] .
(5)
(6)
Here α is a parameter, determined by the logarithmic slope of ξ(r), while
f = d ln D/d ln a, with D(a) being the standard linear growing mode solu-
tion and a -- the cosmological expansion factor (LSS, §11). The solution (5)
assumes Gaussian initial conditions, and the dynamics of clustering is assumed
to be dominated by the gravity of inhomogeneities in a pressure-less fluid of
non-relativistic particles. For all such models, including those with a non-zero
cosmological constant, f ≈ Ω0.6
m (Peebles 1993, §13). If the correlation function
is given by a pure power law, ξ ∝ r−γ, where 0 < γ < 3, the parameter α is
given by
α ≈ 1.2 − 0.65 γ .
(7)
This formalism can be generalized for a non-power law ξ(r) by replacing γ in
Eq. (7) with an effective slope,
γo ≡ − d ln ξ/d ln r ro ,
evaluated at separation r = ro, defined by the condition
ξ(ro) = 1 .
(8)
(9)
JSD tested the equations (5) - (9) against high resolution N-body simulations,
provided by the Virgo consortium (Jenkins et al. 1998). They found an excellent
agreement between the streaming velocity, predicted by their ansatz and the
v12(r), measured from the simulations in the entire dynamical range, 0.1 <
ξ < 103. However, the N-body experiments they used were confined to four
different CDM-like models, considered by Jenkins et al. (1998). As we have
already pointed out in the Introduction, models of this kind fail to reproduce the
observed ξg(r) unless one resorts to a highly contrived, scale- and time-dependent
biasing function (Gaztanaga 1995, Jenkins et al. 1998, Peebles 1999). One of
our objectives here is to test the validity JSD ansatz for v12(r) against a new set
of N-body simulations, which differ significantly from those originally considered
by JSD. Here we use simulations with a more realistic ξ(r), inferred by Baugh &
Gaztanaga (1996) from the measurements of galaxy-galaxy correlations in the
APM survey (see the description below; from now on, we will refer to these
numerical experiments and their initial conditions as APM-like).
2.2. The inflection point
In the gravitational instability theory, newly forming mass clumps are generally
expected to collapse before relaxing to virial equilibrium. If this were so, v12(r)
6
would have to be larger than the Hubble velocity Hr to make v12(r) + Hr
negative. As a consequence, the slope of ξ,
d ln ξ(r)/d ln r = − γ(r) ,
(10)
must increase at separations where ξ(r, t) ≈ 1. This effect was recognized long
ago by Gott & Rees (1975). When the expected "shoulder" was not found in the
observed galaxy-galaxy correlation function, Davis & Peebles (1976) introduced
the so-called pre-virialization conjecture as a way of reducing the size of the
jump in γ(r) (the conjecture involves non-radial motions within the collapsing
clump; see the discussion in LSS, §71 and Peebles 1993, pp. 535 - 541; see
also Villumsen & Davis 1986; Lokas et al. 1996 and Scoccimarro & Frieman
1996). Later observational work showed a clear break in the shape of ξ for
several redshift and angular catalogues, which was early evidence for the linear
to non-linear transition, pointed out by Guzzo and collaborators (see the review
by Guzzo 1997).
The arguments, raised by Peebles and Davis (1976) were qualitative rather
than quantitative. Quarter a century later the precision of N-body simulations as
well as the quality of the observational data have improved dramatically enough
to justify a reexamination of the problem. The actual shape of the correlation
function near ξ = 1 can be investigated with high resolution N-body simulations
like those run by the Virgo Consortium (Jenkins et al. 1998). In all four of the
Virgo models, the slope of ξ(r) exhibits a striking feature. Instead of a shoulder,
or a simple discontinuity in γ(r), however, ξ(r) has an inflection point,
d2ξ(r)/dr2 = 0 ,
(11)
which occurs at a uniquely defined separation r = r∗. At this separation, the
logarithmic slope of ξ reaches a local maximum, d ln ξ/d ln r = −γ∗.
In all
four of the models JSD investigated, the inflection point indeed appears near
the transition ξ = 1, as expected by the earlier speculations, involving the
"shoulder" in ξ. Namely, r∗ is almost identical with the scale of nonlinearity:
r∗ ≈ ro .
(12)
More precisely, a comparison of Figure 1 in JSD with Figure 8 in Jenkins et al.
(1998) gives
ro − r∗ < 0.1 ro
(13)
for all four considered models. Moreover, for all models, studied by JSD, the −γ
vs. r dependence can be described as an S-shaped curve, with a maximum at
r = r∗ ≈ ro, and a minimum at a smaller separation. For r ≥ r∗, the nonlinear
slope (measured from Virgo simulations) follows the linear γ(r), determined by
the initial conditions. The separation r∗ is therefore also the branching point
between the linear and nonlinear γ(r) curves, which are identical for r > r∗
(actually, they differ a little; the small differences between the two curves can
be entirely attributed to sampling errors in the N-body experiments; see Figure
1 in JSD). This property can be used to identify r∗ in noisy simulations, and/or
observations, when the noise in the measured γ(r) curve does not allow a direct
determination of r∗ as the position of the maximum in −γ(r).
7
If the relation (12) is indeed a general property of gravitational clustering, it
can be used as a test of biasing as follows. Suppose the biasing factor is a scale-
independent constant, significantly greater than unity: b ≫ 1. Then ξg ≫ ξ
and the relation (12) will break down. For a power-law correlation function,
ξg(r) = (ro/r)γ = b2ξ(r), and instead of equation (12) we will have
r∗ ≈ ro b−2/γ .
(14)
Since the observed slope is γ ≈ 1.8, for b = 2, the shoulder in the correla-
tion function should appear at a separation smaller than a half of the ro! The
above argument can be generalized to a broader class of models, allowing scale-
dependent bias provided b(r) is a smooth monotonic function and b(ro) is sig-
nificantly greater than unity. Whenever these conditions apply, strong biasing
always implies ro ≫ r∗. Hence, the comparison of these two scales determined
from observations can be used as a diagnostic for biasing.
3. N-body simulations
Initial conditions
3.1.
In this section, we compare our ansatz for v12(r) with the results from P3M
simulations. We use APM-like models for the initial shape of P (k) in Baugh
& Gaztanaga (1996). The box size is 600 h−1 Mpc (or 300 h−1 Mpc) with
2003 (or 1003) dark matter particles. The APM-like models have Gaussian
initial conditions with a power spectrum inferred from correlations in the APM
survey, following the procedure, introduced by Baugh & Gaztanaga 1996. The
power spectrum of the density fluctuations is related to ξ(r) through the usual
Fourier transform,
P (k) = Z ξ(r) exp(ik · r) d3r .
(15)
Following the prescription of Baugh & Gaztanaga, we assume an initial power
spectrum of the form
P (k) =
Ak
(cid:16)1 + (k/0.05h Mpc−1)2(cid:17)1.6 .
(16)
The linear spectrum, given above is designed to evolve into a nonlinear one,
matching the APM measurements of P (k) under the assumption of no bias. The
normalization constant A is directly related to the degree of inhomogeneity of
the mass distribution at the end of the simulation, parametrized by σ8 -- the rms
mass density contrast, measured in spheres of a radius of 8 h−1Mpc. Following
Baugh and Gaztanaga, we choose the constant A in order to ensure that at the
end of the simulation, σ8 = 0.85. The quality of the agreement of the evolved
ξ(r) with the APM data at small separations, where ξ ≫ 1, depends on the
assumed value of Ωm. However, as we show below, for r ≥ 2h−1Mpc and ξ < 3,
this effect is negligible (compare Figures 1 and 3). This range of separations and
clustering amplitudes brackets from both sides the region, on which we focus on
here: the transition between ξ < 1 and ξ > 1. We use simulations with Λ = 0
and two different values of the density parameter: Ωm = 1 and 0.3.
8
3.2. Estimators
Two different estimators for ξ(r) and v12 are used. To construct the first estima-
tor, we cut the simulation box into cubical pixels of size ∆l, placed on a regular
lattice. The density fluctuation amplitude at the ith pixel is
δi =
Ni
hN i
− 1 ,
(17)
where Ni is the particle count in that pixel. The estimator for the two-point
function is then:
(18)
(19)
ξ(1)(r) =
δiδj Wij(r) ,
1
Nr Xi,j
where
Nr = Xi,j
Wij(r)
is the number of pairs of pixels at separation r in the sampled region, and the
window function Wij(r) = 1 if pixels i and j are separated by ~ri − ~rj = r ± ∆r,
and 0 otherwise. For the pairwise velocity we define as vi the average velocity
in the ith pixel at position ~ri. We can then use an equivalent expression for the
first estimator of v12:
where
v(1)(r) = Xi,j
(~vi − ~vj) · rij Uij(r) ,
rij ≡
~ri − ~rj
~ri − ~rj
(20)
(21)
is a unit vector separating pixels i and j and the sum is over all pairs of pixels,
while
Uij ≡
(1 + δi) (1 + δj) Wij(r)
Nr[1 + ξ(r)]
.
(22)
1
The second estimator uses all of the particle pairs rather than the counts in pixel
pairs:
W ′
(23)
ξ(2)(r) =
ij(r) − 1,
hniVr Xi>j
where now the window function W ′
ij(r) = 1 if particles i and j are separated
by ~ri − ~rj = r ± ∆r, hni is the mean particle density and Vr is the volume of
the spherical shell of radius r and thickness ∆r. Typically, hni Vr is estimated
from a random catalogue of particles, drawn from the same sample, because
the effective volume, containing the pairs separated by a distance r ± ∆r might
depend on the geometry. However, our simulations use a large and periodic box,
with no boundaries and high densities (not affected by shot-noise at the scales
of interest). So the above expression gives a good and quick estimator. The
corresponding estimator for the pairwise velocity is
v(2)(r) = Pi>j (~vi − ~vj) · rij W ′
ij(r)
Pi>j W ′
ij(r)
9
(24)
where vi and vj are now individual particle velocities. Note that this estimator
does not depend on the effective volume of the shell, Vr.
Both estimators agree reasonably well in our simulations. The first set is
more useful (faster to run) for large separations, as we can reduce the resolution
of the lattice and have a relatively small number of pixels. The second set is
more adequate (faster to run) for the small separations.
3.3. The correlation function
The evolved, nonlinear correlation functions, measured from simulations are
shown in Figure 1 (top panel). The full squares correspond to the Ωm = 1
model, while the open squares represent Ωm = 0.3. For comparison, we show
the linear correlation function (dashed line), scaled from some "initial" redshift
zo to the present (z = 0), following the linear theory expression for the Einstein-
de Sitter model, ξ(r, 0) = ξ(r, zo)(1+zo)2. Nonlinear effects are more pronounced
in the low density model. This happens because of the well known suppression
of linear growth, which occurs at late times ( z < 1/Ωm) in low density models
and leads to the enhanced clustering on small scales relative to large scales.
Note however, that although the correlation functions differ significantly in
amplitude at separations r < 2h−1Mpc, their slopes γ(r) are almost indistin-
guishable (Figure 1, bottom panel).
The particle resolution (the Nyquist wavelength ∝ N −1/3) of the simulations
used here is significantly lower than the resolution of Virgo simulations, and the
noise in the measured ξ(r) is further amplified by differentiating over r. As a
result, determining the position of the inflection point r∗ directly from the γ(r)
curve alone is difficult. To overcome this problem, we identify r∗ by comparing
the linear and nonlinear γ(r) curves. Taking r∗ to be the separation at which
the nonlinear slope drops below the linear slope in Figure 3, as described earlier,
we get
r∗ ≃ 5 h−1Mpc ≈ ro ,
(25)
in excellent agreement with equation (13). Hence, the equality between r∗ and
ro can probably be considered as a generic outcome of gravitational dynami-
cal evolution in a model where galaxies trace the mass and the initial slope,
d ln ξ/d ln r, is a smooth decreasing function of the separation r. Such a picture
is also known as hierarchical clustering; an obvious additional condition to make
sure that small scale clumps collapse before the large scale ones, is γ > 0, see
e.g. LSS.
3.4. The relative velocity
In this section we describe N-body tests of the JSD model for the relative
motions in pairs of galaxies. We consider two models with APM-like initial
spectra: an Einstein-de Sitter model and an open model with Ωm = 0.3. For
both models, the theoretical predictions for the mean pairwise velocity, based
on equation (5), are plotted in Figure 2 as continuous lines. These predictions
can be compared with N-body measurements, shown as full squares for the
Ωm = 1 model and as open squares for Ωm = 0.3. The agreement between
the theoretical and experimental v12(r) curves shows that our ansatz provides a
good approximation of the N-body results in the entire dynamical range probed
10
Figure 1.
The top panel shows the linear correlation function (dashed
line) and the measured non-linear ξ(r), obtained from the APM-like
simulations with density parameters Ωm = 0.3 (open squares) and
Ωm = 1.0 (full squares). The bottom panel shows the correspond-
ing logarithmic slope, −γ(r) = d ln ξ/d ln r for each of the three curves
from the top panel. The vertical dotted line shows the separation ro,
defined by the condition ξ(ro) = 1 (top) and the separation r∗, at which
the non-linear γ(r) curve crosses the linear one (bottom).
11
Figure 2.
The mean pairwise velocity v12(r), measured from two sets
of APM-like simulations with Ωm = 0.3 (open squares) and Ωm =
1.0 (full squares), are compared with with an approximate analytical
solution of the pair conservation equation (eq. [5]; continuous lines).
for both models. The mean and errors in the Ωm = 1 simulations come from
the mean and dispersion, obtained from five independent realizations of the
APM-like model. For the open model (Ωm = 0.3, open squares) we used only
one realization, but the expected sampling variance is expected to be the same.
Indeed, the initial P (k) is identical in both cases. Moreover, the long-wave tails
of the final power spectra (which determine the size of the sampling error bars)
are also identical because they are not affected by the nonlinear evolution.
4. Comparison with observations
4.1. The correlation function
The measurements of ξg(r), and γ(r) ≡ −d ln ξg/d ln r, obtained from the angu-
lar correlations of galaxy pairs in the APM catalogue (Baugh 1996), are plotted
in Figure 3. The top panel shows the two-point function (points with error bars),
and the linear theory curve, described in §3.3 (dashed line). The intersection of
the two perpendicular dotted lines marks the point (ξg, r) = (1, ro). The bottom
panel of Figure 3 shows the APM γ(r) as a function of the pair separation r.
Note the remarkable similarity between the empirical data and the characteristic
peak in the γ(r) found in the simulations (Figure 1). The intersection of the
two mutually perpendicular, dotted lines in the bottom panel of Figure 3 marks
the crossing between the linear model for γ(r) (dashed line) and the nonlinear
γ(r) curve, determined from the APM catalogue. The crossing occurs at the
12
Figure 3.
The spatial correlation function of APM galaxies (top
panel, symbols with errorbars), compared to the linear theory APM-
like model, described earlier (top panel, dashed line). The bottom panel
shows the corresponding logarithmic slope, γ(r). The intersection of
the two perpendicular dotted lines marks the points where ξ ≃ 1 (top)
and where the non-linear slope crosses the linear one (bottom).
13
separation r ≃ 5 h−1Mpc, and to first approximation this scale could be identi-
fied with r∗. However, a closer inspection of our Figure 3 suggests that, given
the error bars, the actual position the peak could be shifted to the right, to a
somewhat larger separation. Taking into account the error bars in Figure 3 as
well as the uncertainties in the assumed linear theory slope, we obtain
and
r∗ ≃ (6 ± 1) h−1Mpc ,
ro ≃ (5 ± 1) h−1Mpc .
(26)
(27)
The slope at r = r∗ is γ∗ ≃ −1.4.
relation r∗ ≈ ro b−2/γ∗ gives
If we assume the linear bias model, the
b = 1.11 ± 0.22
(28)
at one-sigma statistical significance level.
4.2. The relative velocity
We will now apply the second of the two proposed tests of biasing: the relative
velocity test. We will compare the mean pairwise velocity, predicted by assuming
that the APM galaxies trace the mass with the pairwise velocity, measured
directly from a peculiar velocity -- distance survey.
Figure 4 shows v12(r) curves, predicted by equation (5) (continuous lines) for
three different values of Ωm, from bottom to top Ωm = 1.0, 0.3, 0.1, respectively.
To calculate v12(r), we have used ξ(r), estimated from the APM survey under
the assumptions ξg = ξ and R = 1.
Before making the comparison, we must overcome the following problem.
The survey has a significant depth, with the mean redshift of z ≃ 0.15 while
the measured v12(r) corresponds to the present time (z = 0). To evolve ξ(r, z)
from z ≃ 0.15 to z = 0, we need to make some additional model assumptions.
Gaztanaga (1995) has shown that for this redshift range, the uncertainties in
the details of dynamical evolution of ξ are small.
In particular, choosing an
incorrect value for Ωm can affect ξ at most at the several per cent level (for
Ωm ranging from 1 to 0). Adding this to other possible sources of errors, such
as uncertainties regarding the redshift evolution of the galaxy number density
and sampling and selection fluctuations, Gaztanaga (1995) estimates that the
resulting relative uncertainty in the amplitude of ξg is <∼ 20%. According to
his analysis, the present (z = 0) amplitude of the rms fluctuation of the APM
galaxy counts, measured in spheres of radius of 8 h−1Mpc, is 1.1 <∼ σ8
AP M <∼ 0.9.
To be conservative, for each value of Ωm, we plot the predicted v12(r) curves
for two values of σ8, differing by 20%. The resulting prediction for each value
of Ωm is therefore an area rather than a single v12(r) curve (see Figure 4). The
lower boundary of each shaded area assumes σ8 = 1.1 while the upper boundary
was calculated by assuming σ8 = 0.9. A direct measurement at r = 10 h−1
from the Mark III galaxy peculiar velocity survey (Willick et al. 1997) gives
(Juszkiewicz et al. 2000a)
v12 = −280 ± 60 km/s .
(29)
14
Figure 4.
Predictions for the mean pairwise velocity v12, based on
the assumption that the APM galaxies trace the mass. The shaded
regions correspond to 20% uncertainties in the strength of clustering.
We consider three values of Ωm = 1.0, 0.3, 0.1, top to bottom (as la-
beled). The point with error bars corresponds to a direct measurement
from the Mark III survey (Juszkiewicz et al. 2000a).
It is reassuring that this value, plotted in Figure 4, overlaps with the shaded
area, corresponding to Ωm = 0.3, because it agrees with ranges for Ωm and σ8
obtained by Juszkiewicz et al. from the analysis of the Mark III survey alone.
Their one-sigma constraints are Ωm = 0.35+0.35
−0.25 and σ8 ≥ 0.7, and the analysis
assumes that the correlation function for the mass is well approximated by a
pure power law with γ = 1.75.
From the agreement between the predicted and observed value of the mean
pairwise velocity, we conclude that the Mark III and APM data, taken together,
are consistent with the hypothesis that the APM galaxies trace the mass, b ≈
R ≈ 1, while the density parameter is low, Ωm ≈ 0.3.
4.3. How can biasing affect v12(r) ?
Apart from leading to predictions which are confirmed observationally, the "what
you get is what you see" hypothesis has another important advantage: simplicity.
Once ξ(r) is estimated from observations, a family of v12(r, Ωm) curves can be
calculated directly from equation (5) for any given range of values of the density
parameter. This simplicity will immediately go away if we allow scale-dependent,
stochastic and nonlinear biasing. A frank answer to the question posed in the
15
heading of this subsection would have to be "Only God (of biasing) knows".
Predicting v12(r) without resorting to massive numerical simulations would be
simply impossible. We can get the idea of what is in store by considering only the
leading order term in the perturbative expansion for v12(r) at large separations
(Juszkiewicz et al. 2000b),
v12(r) = − 2
3 Hrf (Ωm) ¯ξgρ(r) ,
(30)
where ¯ξgρ(r) = (3/r3) R r
tion,
0 ξgρ(x)x2dx is the galaxy-mass cross-correlation func-
ξgρ(r) = hδ(0)δg (~r)i ,
(31)
averaged over a sphere of radius r. The function ξgρ describes the cross-correlations
between the mass density and the density of galaxies in the velocity field survey,
which in the case considered here would be the Mark III catalogue.
To make progress in our analysis, we will now generalize the definition of
the stochasticity parameter introduced as a normalized cross-correlation of two
random fields, δ and δg, measured at the same position in space. Instead, we
will consider a cross-correlation of the same two fields measured at two different
positions in space, separated by distance r. Our old equation (2) is replaced by
R(r) =
ξgρ(r)
qξ(r)ξg(r)
.
(32)
Let us make another simplifying assumption, that b2(r) ≡ ξg(r)/ξ(r), as well as
R, are separation-independent. Equation (30) becomes
v12(r) = − 2
3 f (Ωm) Hr R b ¯ξ(r) .
(33)
The expected relative pairwise velocity can now be related to the APM data by
substituting
¯ξ(r) =
bA
0
3
2 r3 Z r
ξg(x)x2dx ,
(34)
2 = ξg(r)/ξ(r) and ξg is the APM galaxy correlation function. In case
where bA
of trouble in predicting the correct v12(r) curve, we now have three essentially
free parameters which can be readjusted. This is only the tip of the iceberg, as
we have ignored nonlinear dynamics as well as the scale-dependence of b and R.
For the linear bias model, the predictions are in clear conflict with the data
unless b is close to unity. After setting R = 1, we get v12(r) ∝ b. Then, if the
biasing factor for spiral galaxies is, as usually assumed b ≈ 1, our predictions for
v12(r) in the linear regime (r >∼ 10 h−1Mpc) should be similar to the unbiased
predictions already plotted in Figure (4). If the biasing factor for the ellipticals
is significantly different, say, b ≈ 2, the elliptical subsample of the Mark III
data should give estimates of v12 which differ from the estimates from the spiral
sample by the same factor of two. Meanwhile the estimates from the appropriate
subsamples in the real data are indistinguishable (Juszkiewicz et al. 2000a).
Hence, just as in case of the shape of the APM correlation function, considered
above, the deterministic linear biasing model is inconsistent with observations.
16
We can summarize the last two subsections as follows. The prediction for
v12(r), based on the assumption that the APM galaxies trace the mass passes
our test as it agrees with the velocity, estimated from the Mark III data. The
simplest prescription of biasing fails the test. More complicated prescriptions
can probably be made to pass, which is not surprising, given the number of free
parameters.
5. Summary and discussion
A quarter of a century ago, Gott and Rees predicted that gravity should leave
its mark on the shape of the galaxy autocorrelation function: a "shoulder", or
steepening of the slope of the correlation function should appear near the sep-
aration where ξ passes through unity. At the time, biasing was unheard of,
and Gott and Rees (1975) assumed ξ = ξg. Recently, in another context, JSD
have studied the the ξ = 1 boundary in the evolution of the mass correlation
function, using results from Virgo simulations. They found that the "shoulder"
is actually an inflection point, occurring at a well defined separation r∗. In all
four CDM-like models they studied, the nonlinear transition looked strikingly
similar: the inflection occurred at almost the same separation as that of the
nonlinear transition: r∗ ≈ ro, where ro corresponds to ξ = 1. Here we have
tested the degree of universality of their results by widening the range of mod-
els considered. Our additional objective was to study the range of validity of
an approximate solution of the pair conservation equation, proposed by JSD
to study the nonlinear evolution of the relative velocity of particle pairs at a
fixed separation, v12(r). We used N-body simulations, with APM-like initial
conditions, with two different values of the density parameter: Ωm = 1 and 0.3.
The APM-like initial power spectra differ significantly from all of the CDM-
like spectra, considered earlier by JSD. Moreover the spectra of the latter kind
appear as more realistic to us because they can reproduce observations without
resorting to scale-dependent biasing. Our APM-like simulations are in excel-
lent agreement with earlier results, confirming the validity of the JSD ansatz for
v12(r) and the conjecture that the appearance of the shoulder in the correlation
function near the ξ = 1 transition is a feature of gravitational dynamics rather
than a peculiarity of a particular set of initial conditions.
Using these results, we proposed two tests of the hypothesis that galaxies
trace mass. The first of the tests is based on an obvious idea, that if ξ(r) =
ξg(r), the galaxy correlation function near ξg = 1 should exhibit properties
similar to those of the matter correlation function. We examined the behavior
of the correlation function, derived from the APM catalogue and found exactly
the same features we knew earlier from N-body simulations, in particular the
agreement between the two characteristic scales, r∗ ≈ ro. It is difficult to imagine
how such an agreement could happen by a mere coincidence, which would have to
be the case if ξg is unrelated to ξ. The agreement between the two characteristic
scales can be used to constrain the linear biasing factor for the APM catalogue
to be within 20% of unity. This constraint agrees with an earlier limit, obtained
from measurements of the three-point correlation function from the APM survey
(Gaztanaga 1994, Frieman & Gaztanaga 1999).
17
The second test confronts the v12, predicted by assuming that the APM
galaxies are unbiased tracers of mass with direct measurements of v12. The
results are again consistent with b ≈ 1 and a low density parameter, Ωm ≈ 0.3,
in agreement with the limits obtained from the velocity data alone (Juszkiewicz
et al. 2000a).
We are impressed how well the observations are reproduced by the simple
calculations based on the assumption that galaxies follow the mass distribution,
at least on large (weakly non-linear) scales. We are unable to constrain biasing
models with a large number of free parameters, but their predictive power is
questionable and one may ask: are such models falsifiable and therefore worth
constraining?
Our results are by no means final, they are also less rigorous than one could
wish because we are limited by the accuracy of the present observational data.
New generation of catalogues promise a dramatic improvement on this front in
the near future (for an excellent collection of reports on the state of the art in
this field, see Colombi et al. 1998).
6. Acknowledgements
One of the co-authors (RJ) who was more lucky than the other and who at-
tended this excellent Meeting would like to thank Jim Fry for making it happen.
We also thank Carlton Baugh for providing us with his APM-like simulations as
well as his estimate of ξg(r), based on the APM survey. RJ thanks Ruth Durrer
for important discussions regarding the inflection point in ξ(r) and for her hospi-
tality at the University of Geneva. This work was supported by a collaborative
grant between the Polish Academy of Science and the Spanish Consejo Superior
de Investigaciones Cientificas. We also acknowledge support by grants from the
Polish Government (KBN grant No. 2.P03D.01719), the Swiss Tomalla Foun-
dation, and from IEEC/CSIC and DGES(MEC) (Spain), project PB96-0925.
References
Binney, J., & Merrifield, M., 1998, Galactic Astronomy, Princeton University
Press, Princeton
Baugh, C.M., 1996, MNRAS, 282, 1413
Baugh, C.M., & Gaztanaga, E., 1996, MNRAS, 280, 37
Bernardeau, F., 1994a, A&A, 291, 697
Bernardeau, F., 1994b, ApJ, 433, 1
Blanton, M., Cen, R., Ostriker, J.P., Strauss, M.A., 2000, ApJ, 531, 1
Colombi, S., Mellier, Y., & Raban, B., eds., 1998, Wide Field Surveys in Cos-
mology, Editions Frontieres, Paris.
Davis, M. & Peebles, P.J.E., 1977 ApJS, 34, 425
18
Davis, M. & Peebles, P.J.E., 1983, ApJ, 267, 465
Davis, M., Efstathiou, E., Frenk, C.S., White, C.D.M., 1985, ApJ, 292, 371
Dekel, A., & Lahav, O., 1999, ApJ, 520, 24
Feldman, H.A., Frieman, J.A., Fry, J.N. & Scoccimarro, R., 2000, preprint,
astro-ph/0010205
Fischer, P., et al. (the SDSS Collaboration), 1999, preprint, astro-ph/9912119
Fisher, K.B., Davis, M., Strauss, M., Yahil A., Huchra, J. 1994, MNRAS, 267,
927
Frieman, J.A., Gaztanaga, E., 1999, ApJ, 521, L83
Fry, J. 1996, ApJ, 461, L65
Gaztanaga, E., 1994, MNRAS268, 913 (1994)
Gaztanaga, E., Frieman, J.A., 1994, ApJ, 437, L13
Gaztanaga, E. 1995, ApJ, 454, 561
Gott, J.R., & Rees, M.J., 1975, A&A, 45, 365
Guzzo, L., 1997, New Astronomy, 2, 517
Hamilton, A.J.S., 1998, in Hamilton D., ed., The Evolving Universe, Kluwer,
Dordrecht, p. 185
Hamilton, A.J.S., Tegmark, M., & Padmanabhan, N., 2000, preprint (astro-
ph/0004334)
Hamilton, A.J.S., Tegmark, M., 2000, preprint, astro-ph/0008392
Jenkins, A. et al. (The Virgo Consortium), 1998, ApJ, 499, 20
Juszkiewicz, R., Fisher, K., & Szapudi, I., 1998, ApJ, 504, L1
Juszkiewicz, R., Springel, V., & Durrer, R., 1999, ApJ, 518, L25 (JSD)
Juszkiewicz, R., Ferreira, P.G., Feldman, H.A., Jaffe, A.H., & Davis, M., 2000,
Science, 287, 109
Juszkiewicz, R., Durrer, R., Ferreira, P., 2000, in "Energy densities in the Uni-
verse", Proc. Recontres de Moriond, in press
Lokas, E., Juszkiewicz, R., Bouchet, F.R., & Hivon, E., 1996, ApJ, 467, 1
Peebles, P.J.E., 1980, The Large -- Scale Structure of the Universe, Princeton
University Press, Princeton (LSS)
Peebles, P.J.E., 1993), Principles of Physical Cosmology, Princeton University
Press, Princeton
19
Peebles, P.J.E., 1999, in Clustering at High Redshift, eds. A. Mazure & O. Le
Fevre (astro-ph/9910234)
Rees, M.J., 1999, preprint, astro-ph/9912373
Saunders, W., 2000, preprint, astro-ph/0001117
Seljak, U., 2000, preprint, astro-ph/0004086
Scoccimarro, R., & Frieman, J. 1996, ApJ, 473, 620
Scoccimarro, R., Sheth, R., Hui, L., & Jain, B., 2000, ApJ, in preparation
Somerville, R.S., et al., 2000, MNRAS, 320, 289
Strauss, M., Willick, J., 1995, Physics Reports, 26, 271
Van Waerbeke et al., 2001, preprint, astro-ph/0101511
Villumsen, J., & Davis, M., 1986, ApJ308, 499
Vittorio, N., Juszkiewicz, R., & Davis, M, 1986, Nature, 323, 132
Willick, J., et al., (1997), ApJS, 109, 333
20
|
astro-ph/9901267 | 1 | 9901 | 1999-01-20T14:32:41 | A Deep 12 Micron Survey with ISO | [
"astro-ph"
] | We present the first results of a study of faint 12 micron sources detected with the Infrared Space Observatory (ISO) in four deep, high galactic latitude fields. The sample includes 50 such sources in an area of 0.1 square degrees down to a 5 sigma flux limit of about 500 microJy. From optical identifications based on the US Naval Observatory (USNO) catalogue and analysis of the optical/IR colours and Digital Sky Survey (DSS) images, we conclude that 37 of these objects are galaxies and 13 are stars. We derive galaxy number counts and compare them with existing IRAS counts at 12 microns, and with models of the 12 micron source population. In particular, we find evidence for significant evolution in the galaxy population, with the no-evolution case excluded at the 3.5 sigma level. The stellar population is well matched by existing models. Two of the objects detected at 12 micron are associated with known galaxies. One of these is an IRAS galaxy at z=0.11 with a luminosity of 10^11 L_sun. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
1(11.06.1; 11.09.2; 13.09.1
ASTRONOMY
AND
ASTROPHYSICS
A Deep 12 Micron Survey with ISO ⋆
D.L. Clements1, F-X. Desert1,2, A. Franceschini3, W.T. Reach1,4, A.C. Baker5 J.K. Davies6 and C.
Cesarsky5
1Institut d'Astrophysique Spatiale, Batiment 121, Universite Paris XI, F-91405 Orsay CEDEX, France
2Laboratoire d'Astrophysique, Observatoire de Grenoble, BP 53, 414 rue de la piscine, 38041 Grenoble CEDEX 9, France
3Dipartimento di Astronomia, Universita' di Padova, Padova, Italy
4IPAC, Caltech, MS 100-22, Pasadena, CA 91125, USA
5Service d'Astrophysique, Orme des Merisiers, Bat. 709, CEA/Saclay, F91191 Gif-sur-Yvette CEDEX, France
6JACH, 660 N. Aokoku Place, University Park, Hawaii, HI 96720, USA
Received; accepted
Abstract. We present the first results of a study of faint
12µm sources detected with the Infrared Space Observa-
tory (ISO) in four deep, high galactic latitude fields. The
sample includes 50 such sources in an area of 0.1 square
degrees down to a 5σ flux limit of ∼ 500µJy. From op-
tical identifications based on the US Naval Observatory
(USNO) catalogue and analysis of the optical/IR colours
and Digital Sky Survey (DSS) images, we conclude that 37
of these objects are galaxies and 13 are stars. We derive
galaxy number counts and compare them with existing
IRAS counts at 12µm, and with models of the 12µm source
population. In particular, we find evidence for significant
evolution in the galaxy population, with the no-evolution
case excluded at the 3.5σ level. The stellar population is
well matched by existing models. Two of the objects de-
tected at 12µm are associated with known galaxies. One
of these is an IRAS galaxy at z=0.11 with a luminosity of
1011L⊙.
Key words:
stars:infrared
galaxies:evolution,
galaxies:infrared,
1. Introduction
Whenever a radically new wavelength or sensitivity regime
is opened in astronomy, new classes of object and unex-
plained phenomena are discovered. The IRAS satellite, in
opening the mid- to far-IR sky, revealed that the bolo-
metric luminosities of many galaxies are dominated by
offprint
requests
Institut
Send
d'Astrophysique Spatiale, Batiment 121, Universite Paris
XI, F-91405 ORSAY Cedex, France
to: D.L. Clements,
⋆ Based on observations with ISO, an ESA project with in-
struments funded by ESA Member States (especially the PI
countries: France, Germany, the Netherlands and the United
Kingdon) and with the participation of ISAS and NASA
this spectral region. This has raised questions concern-
ing the evolution of galaxies in the mid- to far-IR, the
role of dust-extinction in the formation of galaxies, the
relationship between quasars and galaxies, and the nature
of galaxy formation itself. Limited as they were to fluxes
not much smaller than 1Jy, the IRAS surveys were con-
strained to the fairly local universe for the vast majority
of the detected objects. The evolutionary properties of IR
galaxies, both the normal galaxy population and the un-
usual ultra- and hyper-luminous objects (see e.g. Clements
et al. 1996a), are thus still mostly unknown.
Recent work in the visible and near-IR has had a ma-
jor impact on our understanding of the star formation
history of galaxies (eg. Steidel et al., 1996, Cowie et al.,
1994, Giavalisco et al., 1996, Lilly et al., 1996). It appears
that the star formation rate in the universe peaked at
around z=1 and has been declining since (Madau et al.
1998). Many of the objects studied in deep, high redshift
fields appear to be distorted, and are possibly undergoing
tidal interaction or merging (Abraham et al., 1996). It is
well -- known that tidal interactions and mergers in the lo-
cal universe produce significant amounts of star formation
(Joseph & Wright, 1985, Clements et al., 1996b, Lawrence
et al., 1989), and that these are usually associated with a
significant luminosity in the mid- to far-IR. The dust re-
sponsible for this emission is heated by the star formation
process, which it also obscures. We must thus consider
that the view of the universe obtained in the visible and
near-IR, corresponding to the rest-frame optical and near-
UV of many of the objects observed, may well be biased
by such obscuring material. The question of how much
of the star-formation in the universe is obscured by dust
thus becomes important. This issue can only be properly
addressed by selecting objects in the mid- or far-IR which
are less affected by such obscuration.
Previous work in the mid- and far-IR used data from
IRAS, with all-sky sensitivities of ∼0.1 Jy in the mid-IR
bands (12 and 25 µm), and ∼ 0.3 -- 1 Jy at 60 and 100
2
D.L. Clements et al.
µm. These typically allow the detection of galaxies out
to z=0.2, though a few exceptional objects, such as the
gravitationally lensed z=2.286 galaxy IRAS10214+4724
(Rowan-Robinson et al., 1991, Serjeant et al., 1995), have
also been detected.
Most evolutionary studies with IRAS have concen-
trated on the 60µm waveband (Sanders et al., 1990, Hack-
ing & Houck, 1987, Bertin et al., 1997). This work has
found evidence for strong evolution in the 60µm popula-
tion, at rates similar to those of optically selected AGN,
but the nature of this evolution is still unclear, and it is
difficult to extrapolate to higher redshift with any confi-
dence.
At mid-IR wavelengths, the IRAS mission has pro-
duced both large-area surveys of fairly nearby objects (eg.
Rush et al., 1993), and small-area, deep surveys in re-
peatedly scanned regions (eg. Hacking & Houck, 1987).
The former surveys do not probe sufficiently deeply into
the universe to be able to say much about galaxy evolu-
tion, but they do have the advantage that plentiful data
exists for the nearby galaxy samples they produce. The
latter surveys are plagued by stellar contamination. The
vast majority of the 12µm objects in the survey of Hacking
et al., for example, are stars -- there are only five galaxies
in their entire survey.
The Infrared Space Observatory (ISO, Kessler et al.
1996) provides a major improvement to our observational
capabilities beyond those of the IRAS satellite. For ob-
servations in the mid-IR, the ISOCAM instrument (Ce-
sarsky et al. 1996) allows us to reach flux limits ∼100
times fainter than those achieved by IRAS while observ-
ing fairly large areas (∼0.1 sq. degree) in integration times
of only a few hours. We can thus probe flux regimes that
were previously impossible to study.
This paper presents the results of a survey of four high
galactic latitude fields using the LW10 (12µm) filter, which
was specifically designed to match the 12µm filter on the
IRAS satellite. The present results can thus be compared
to existing IRAS data with minimal model-dependent K-
corrections. This survey is much deeper than any based
on IRAS data, and is sufficiently deep to detect distant
galaxies. The survey region is also fairly small and at high
galactic latitude, so that stellar contamination should not
be a major problem.
There are of course other studies in the mid-IR un-
derway using the ISO satellite. These include the DEEP
and ULTRADEEP surveys (Elbaz et al., in preparation),
the ELAIS survey (Oliver et al., in preparation) and the
ISOHDF project (Oliver et al., 1997; Desert et al., 1998
(Paper I); Aussel et al. 1998). All of these programmes use
the LW2 6.7µm and/or LW3 15µm filter on ISO. Only the
ISOHDF results have been published to date. At 15µm
these observations reach a flux limit of ∼0.1 mJy, about
5 times deeper than the observations discussed here, but
cover only 1/24th of the area. There are also deep sur-
veys at longer wavelengths using the PHOT instrument
at 175µm (Kawara et al., 1998, and Puget et al., 1998).
Future missions will also be probing this part of the elec-
tromagnetic spectrum. The first of these will be the WIRE
mission (Fang et al., 1996) which will obtain a large area,
deep survey at 12 and 25µm. The SIRTF project (Werner
& Bicay, 1997) and IRIS satellite (Okuda, 1998) will also
be used for deep number counts, and should be able to
probe significantly deeper than ISO. Finally the planned
NGST (Stockman & Mather, 1997) will provide incompa-
rable performance in this cosmologically interesting wave-
band. The present work provides the first results of the
exploration of the distant universe at these mid-IR wave-
lengths, and can provide a guideline for future missions,
useful for their planning and preparation.
The paper is organised as follows. Section 2 describes
the observations, data reduction and calibration. Section 3
provides details on identifications of the 12µm source pop-
ulation, star-galaxy separation, and on individual source
properties. Section 4 discusses number counts, compari-
son with statistics at other wavelengths, and with model
predictions. Section 5 summarises our conclusions.
2. Observations and Data Reduction
trails
The observations presented here were part of a
survey of cometary dust
(ISO project name
JDAVIES/JKDTRAIL). The original goal was to observe
the width and structures of the cometary trails, which are
produced by large particles, ejected from the comet into
independent heliocentric orbits but with very similar or-
bital elements and very small radiation pressure effects.
The comet 7P/Pons-Winnecke was selected because of its
similarity to other comets with dust trails, the detection of
a dust trail by IRAS (Sykes & Walker, 1992) and its suit-
ability for observation by ISO. Four fields were imaged,
each field being a raster map centred on the ephemeris
prediction for particles with the same orbital elements as
the nucleus except for the mean anomaly of the orbit,
which was shifted by +1◦ (ahead) and −0.5◦, −1◦, and
−2◦ (behind). Each raster was 11 by 7 pointings, with
a spacing of 60′′ by 48′′. The pixel field of view was 6′′,
so that there was substantial overlap between individual
frames of the 32 by 32 pixel ISOCAM LW array (Cesarsky
et al., 1996) to ensure good flatfielding and high obser-
vational redundancy. A typical position on the sky was
visited 12 times during the raster, each time with a dif-
ferent pixel of the array. The rasters were rotated such
that the predicted cometary trail would run parallel to
the short axis of the raster. Unfortunately for the cometry
trails programme, the observations took place on 17 Au-
gust 1997, one day later than assumed for the ephemeris
calculations and therefore the trail is predicted to run hor-
izontally across the very bottom edge of the image. Based
on the results of observations of the trail of another comet
(P/Kopff) from the same observing program (Davies et
al., 1997), we expect that the trail would occupy at most
Deep 12 Micron Survey with ISO
3
the lower 1′ of the image and that it would be relatively
smooth. The presence of the dust trail in the field should
not have any effect on the observations presented here,
though an excess of sources in the bottom of the raster
could potentially be related to structures in the dust trail.
No such excess is seen. All observations presented here are
in the LW10 filter, which is very similar to the IRAS 12
µm filter in wavelength-dependent response. Since these
fields are at high galactic lattitude (>50 degrees), and, in
the absence of cometary trails, are effectively blank fields,
they become ideal for a deep survey of the extragalctic
12µm source population. The positions of the fields are
given in Table 1.
The data reduction process is described in detail in
Paper 1. Basically, for each AOT file (total 4) the raw
data (CISP format) and pointing history (IIPH format)
are read and merged. The raw data cube (typically 1244
readouts of 32 by 32 pixels) is deglitched for fast and
slow cosmic ray impacts. A transient correction is ap-
plied to recover the linearity and the nominal flux response
of the camera. A triple -- beam method is then applied to
the processed data cube, in order to find the best (ON-
(OFF1+OFF2)/2) differential value of the sky brightness
for each pixel and each raster position, where OFF1 and
OFF2 refer to the previous and next raster position value
for the same camera pixel. The resulting low -- level reduced
cube is then simply flat -- fielded (there is no need for a dark
correction because we perform a differential measurement
on the sky). The flat -- fielded reduced cube along with an
error cube is made up of 77 values for each of the 32 by 32
pixels of the camera. The 2 cubes are then projected onto
the sky using an average effective position for ISO dur-
ing each raster position, with a 1/σ2 optimal weighting.
A noise map is thus calculated as well as a sky differen-
tial map. Any given source will leave 2 negative half -- flux
ghosts 60 arcseconds away on both sides along the raster
scan direction. This is a trade -- off in order to beat the 1/f
noise regime that is reached by the camera for long in-
tegrations per position. On the final map, shown in Fig.
1, we search for point sources with a Gaussian fitting al-
gorithm that uses the noise map for weighting the pixels
and deducing the noise of the final flux measurement. A
FWHM of 9 arcseconds was used. The final internal flux
is converted to µJy by using the ISOCAM cookbook value
(Cesarsky et al., 1994). The present understanding of ISO-
CAM calibration is that no additional factor should be
applied to the pre-flight sensitivity estimates for the de-
termination of surface brightness.
We have devised a scheme to assess the reproducibility
of the sources, in order to test for false sources that would
be due to undetected glitches. This is described in detail in
Paper I, but consists of breaking the observations of each
object into a number of independent subsurveys. Sources
are deemed reliable if they are detectable, with suitably re-
duced significance, in each of these subsurveys. Of the 193
sources above the 3σ limit only 7 fail the reproducibility
Field
Field 1
Field 2
Field 3
Field 4
RA
03 05 30
03 00 40
03 09 50
03 04 00
l
190.327
190.813
189.979
190.477
Table 1. Details of 12µm Survey Fields
Dec
-09 35 00
-10 42 00
-08 37 00
-09 56 00
b
-53.895
-55.466
-52.433
-54.361
RA and Dec are in J2000. All fields have high galactic latitude
>50 degrees, making them ideal for cosmological studies.
test (4%) and these are not considered in the following.
Visual screening helped in removing a further 38 resid-
ual companions of strong sources due to imperfect fitting.
Visual screening also showed that two sources, F1 0 and
F2 0, were significantly extended at 12µm. We thus use
aperture photometry to obtain an accurate flux for these
objects. The aperture used had a diameter of 20 arcsecs.
Simulation of the expected PSF from ISO after the
same processing reveals that part of the flux is missed
by the optimised Gaussian fitting. We therefore correct
the fluxes and errors by a factor of 1.52 determined from
this modelling. The final absolute photometry should be
in error by no more than an estimated 30%. The fluxes
are given at the nominal wavelength of 12 µm (i.e. an
additional correction of 1.04 = 12/11.5 is applied since
the nominal ISOCAM calibration is for a wavelength of
11.5µm), in order to facilitate the comparison with previ-
ous IRAS observations. This assumes a flat spectrum in
νFν as was used for IRAS calibration. The flux prediction
for known stars should thus be colour corrected, since they
have a Rayleigh-Jeans spectral index, in order for com-
parison to ISOCAM measurements: the real flux at 12µm
should be divided by 0.902. An additional factor comes
from the fact that the PSF is smaller for stellar sources
(which are dominated by the short wavelength part of the
broad filter) than for the assumed extragalactic sources
which have a broader spectrum. Thus the real flux should
also be multiplied by a supplementary factor of 1.13 (see
Section 3.1 for this a priori calibration and the compari-
son with flux measurements of known stars in the fields).
The basic calibration of ISOCAM, before the corrections
for point sources are applied, can be checked by comparing
the integrated surface brightness of these fields with val-
ues interpolated from the DIRBE experiment on COBE
(Hauser et al. 1997a). The ISO surface brightnesses agree
with the DIRBE values to better than 5%.
The sensitivity that is achieved in the central area
of the fields is about 1σ = 100µJy. This was achieved
with a total integration time of 4 minutes for each cam-
era field -- of -- view. Astrometry was assessed by matching
ISO sources to bright stars in the fields. We estimate the
astrometric accuracy to be ∼6" (2 σ).
4
D.L. Clements et al.
Fig. 1. Maps of the final reduced ISOCAM data for the four 12µm survey fields
Deep 12 Micron Survey with ISO
5
3. The 12µm Source Population
A total of 186 candidate 12µm sources are found in the
survey to a 3σ flux limit of ∼ 300µJy. Visual inspection
then removes 38 of these sources as fragments of brighter
sources incorrectly identified as separate objects, giving a
master list of 148 objects. For the remainder of this pa-
per we shall restrict ourselves to discussion of only those
sources detected at 5σ sensitivity or above in this master
list. This is for several reasons. Firstly, a number of uncer-
tainties remain in the identification of the weakest sources.
These are the sources most likely to be affected by the
remnants of subtracted glitches or by weak, undetected
glitches. Further examination of the detailed time histo-
ries and reproducibilities of these sources is underway, and
a full catalogue reaching to the faintest flux limits can then
be constructed. Secondly, the problems of Malmquist bias
(Oliver, 1995) are most easily controlled in catalogues de-
tected with significances ≥ 5σ. A source list using a 5σ
detection threshold is thus best suited to our examination
of the 12µm source counts. 50 objects are detected at 5σ
or greater significance. Details of these objects are given
in Table 2, and they are discussed further in the following
sections.
3.1. Optical Identifications and Star-Galaxy Separation
Comparison of the 12µm ISOCAM images with Digital
Sky Survey (DSS)1 images shows that a number of the
sources are associated with bright stars. Before we are
able to analyse the galaxy component of the 12µm source
population, these and any other contaminating stars must
be identified and removed. This was achieved by using the
US Naval Observatory (USNO) all-sky photometric cata-
logue (Monet et al., 1996). The database was searched for
all optical objects within 12 arcseconds of each ISO po-
sition. 12 sources were immediately identified with HST
Guide Star Catalogue (GSC) stars, though inspection of
the DSS images shows that one of these is in fact a galaxy
(03 01 06.16 -10 44 23.6, the GSC 'star' 5290 640). B and
R band photometry was extracted for 29 of the 32 opti-
cally identified objects -- three of the GSC stars were too
bright to allow photometry from the B survey plates used
by the USNO catalogue. These magnitudes were then cor-
rected for the estimated galactic extinction. A comparison
of the final FB/FR and F12/FR flux ratios was then made.
Figure 2 shows the optical/ISO colour-colour diagram, to-
1 Based on photographic data of the National Geographic
Society -- Palomar Observatory Sky Survey (NGS-POSS) ob-
tained using the Oschin Telescope on Palomar Mountain. The
NGS-POSS was funded by a grant from the National Geo-
graphic Society to the California Institute of Technology. The
plates were processed into the present compressed digital form
with their permission. The Digitized Sky Survey was produced
at the Space Telescope Science Institute under US Government
grant NAG W-2166.
gether with a Black Body colour track. Simple stars, with-
out associated dust or stellar companions, should lie on or
near to this colour track. As can be seen, almost all of
the GSC stars and several other objects lie near to the
Black Body line. This allows us to remove all those stars
that have not been identified in the GSC. Three such ob-
jects are removed. One star (F3 9: 03 09 42.14 -08 35 44.6)
seems to be anomalously blue (B-R = -0.8 from the USNO
catalogue). However, this object and another bright star
(F3 0: 03 09 42.6 -08 35 33.4) are so close to one another
that accurate photographic photometry is likely to be dif-
ficult, resulting in the anomalous colours. These objects
are removed from further analysis.
Of the 32 optically identified 12µm sources we thus
conclude that 13 are stars and that the remaining 19 are
optically identified galaxies. 18 sources, all probably galax-
ies, thus remain without optical identifications to the lim-
its of the USNO-A catalogue ie. around 20th magnitude
in B and R.
The colour-colour plot also allows us to check the cal-
ibration for the 12µm survey. We can use the B-R colours
to provide a rough spectral type for all stars in the survey.
This can then be cross-referenced to the surface temper-
ature of that stellar type. The 12µm flux can then be
extrapolated from the R band flux, assuming a simple
Black Body spectrum. This approach suggests that the
flux calibration is accurate at the ∼20% level (see Ta-
ble 3). We have also checked these results using detailed
spectral energy distributions (SEDs) based on the Kurucz
stellar atmosphere codes instead of a simple Black Body
extrapolation, and arrive at very similar conclusions. The
main source of uncertainly here is the treatment of un-
dersampled unresolved sources in the ISOCAM reduction
systems. As more data becomes available on the details
of the ISOCAM PSF, this systematic uncertainty will be
reduced. There is also the possibility that one or more of
the stars in the survey have genuine IR excesses. Ground-
based near- to mid-IR photometry will be required to con-
firm this.
We are then able to remove all 13 stars from the 12µm
source lists generated in this survey, and can thus examine
the statistics of faint galaxies at 12µm.
3.2. Individual Sources
We discuss here individual sources of note in this survey.
IRAS 03031-0943 This lies at 03 05 36.4 -09 31
27.0 (J2000) and is an IRAS source identified with a
B=18.6 galaxy at z=0.112 (Clements et al., 1996a). It is
associated with object F1 0 in the present survey. This
galaxy has IRAS fluxes of 0.85 and 0.51 Jy at 100 and
60µm respectively, and limits of 0.15 and 0.095 Jy at 25
and 12µm, consistent with the measured ISO 12µm flux
of 12.1±0.2 mJy. This galaxy has a 60µm luminosity of
1011L⊙ (H0=100kms−1Mpc−1, q0=0.5) which places it
among the high luminosity IRAS galaxies but at lower lu-
6
D.L. Clements et al.
Fig. 2. Optical-IR Colour-Colour Diagram
Triangles are anonymous 12µm sources, stars are objects iden-
tified with GSC stars, the line shows the locus for Black Bodies,
with temperature decreasing from 10000 K in the lower left.
Points on the line lie at 1000K intervals. Note the concentration
of stars on the Black Body line, and the majority of sources ly-
ing well away from the line. This allows star-galaxy separation.
minosity than the ultraluminous class (see eg. Sanders &
Mirabel, 1996). Its optical spectrum contains strong emis-
sion from the Hα-NII blend and SII, but the redshift mea-
surement spectrum is of too low a resolution to provide
any emission line diagnostics (Clements, private commu-
nication). We thus do not know what sort of power source
is energetically dominant in this object -- starburst or
AGN.
NPM1G-10.0117 This lies at 03 01 06.2 -10 44 24
(J2000) and is a B=16.63 galaxy used in a proper-motion
survey (Klemola et al., 1987). It is associated with object
F2 0 in this survey, and has a 12µm flux of 9.10±0.26 mJy.
It is also identified with HST Guide Star GSC 5290 640,
but is clearly a galaxy in the Digitised Sky Survey images.
Little else is currently known about it.
Altogether, we can say relatively little about the galax-
ies identified so far in this survey since the identification
programme has only just started. Nevertheless, it is a use-
ful check of the processing to note that the only IRAS
galaxy within the survey region has been detected by ISO-
CAM.
4. The 12 µm Number Counts
Integral number counts from a survey with homogeneous
sensitivity are calculated by summing up the number of
sources to a given flux limit, and then dividing by the
survey area:
N (> S) = X
f lux>S
1
Ω
(1)
Fig. 3. Areal Coverage Plot for 12µm Survey
where Ω is the area of the survey. However,
in our
case the noise in the survey is inhomogeneous (it has a
bowl-like shape) since the border pixels were observed
with smaller integration times. We thus have to make a
correction to equation 1 to account for this. If we define
η(σ) = Ω/Ω(≤ s), where Ω(≤ s) is the area where the 1σ
sensitivity is better than s, then the corrected number
counts are given by:
N (> S) = X
f lux>S
1
Ω
×
1
η(S/n)
(2)
where n is the detection threshold of the survey. In the
present paper we consider only those sources detected at
> 5σ confidence, so n=5. We plot the area coverage, η, in
Fig. 3.
A correction must also be applied to account for
Malmquist bias (Oliver, 1995). This bias arises when look-
ing at number counts for a population with rapidly in-
creasing numbers at fainter fluxes, as is the case for our
12µm galaxy sample. In the presence of observational
noise, some galaxies close to, but below, the flux limit
will be scattered above the flux limit by noise and will ap-
pear in the final catalogue. Similarly some galaxies close
to but above the flux limit will be scattered out of the
catalogue. However, since there are many more galaxies
at fainter fluxes, more galaxies will be scattered above
the flux limit than below it. Number counts that are un-
corrected for this bias thus show a steep rise in counts
towards the faintest flux levels. In the case of Gaussian
noise and a Euclidean count slope, which approximates to
the present case, Murdoch et al (1973) tabulated the ef-
fects of this bias to a detection level of 5σ, allowing for the
observed fluxes to be corrected. Oliver (1995) provides a
numerical version of this correction which we apply here.
For observations probing below the 5σ limit this simple
correction cannot be applied, and a more complex Monte
Carlo approach must be adopted (eg. Bertin et al. 1997).
Deep 12 Micron Survey with ISO
7
Fig. 4. 12µm Integral Galaxy Counts from ISO
This plot presents the current observational state of knowledge on galaxy counts at 12µm. Since the errors in integral-counts
plots are correlated, we show the areas allowed within the ±1σ error bars as shaded regions. The cross-hatched region indicates
the deep IRAS counts from Hacking & Houck (1987) including stars. Triangles indicate the galaxy content of this survey (lower)
and the Gregorich et al. survey (1995) (upper). The solid region in the bottom left shows the counts from the extended IRAS
galaxy survey by Rush et al. (1993) using the all-sky IRAS Point source Catalogue, with the area of this survey taken from
Oliver et al. (1997). The single-hatched region shows the results of the present work, reaching fainter magnitudes than previous
12µm studies. The 13 stars among the 50 objects detected at > 5σ are not included in these count statistics. The triangle in
the top right shows the ISOHDF counts at 15µm (Oliver et al. 1997) -- note that no correction has been applied to this data to
convert from 15 to 12 µm. The solid line is the strong evolution model which is discussed in the text.
Figure 4 shows the Malmquist-bias corrected integral
number count plot from the present work and from 12µm
IRAS surveys, along with some other information. The
first thing to notice in this diagram is that we have reached
flux limits almost 100 times fainter than the deepest IRAS
number counts at these wavelengths. We are thus able to
see much deeper into the universe than the IRAS surveys
and can provide considerably more powerful sampling of
the 12µm galaxy population.
Secondly, our survey is the first flux limited 12µm sur-
vey to be dominated by galaxies rather than stars. The
deepest IRAS sample (Hacking & Houck 1987) included
∼50 objects of which only 5 were galaxies. As discussed
above, the present survey contains 50 objects above the
5σ flux limit, of which only 13 are stars.
The integral counts of stellar identifications in the 12
µm survey are shown in Fig. 5. We find good agreement
with model counts by Franceschini et al. (1991) based on
the Bahcall and Soneira galactic model and on a stellar
luminosity function scaled from the V band to λ = 12 µm
according to Hacking & Houck (1987). This agreement
suggests that no major new stellar component is emerging
at faint fluxes with respect to those detected in the optical.
We have so far shown the integral number counts for
the galaxies in our survey. A more statistically mean-
ingful way to compare observed and theoretical num-
ber counts is to examine them in a differential form, i.e.
8
D.L. Clements et al.
co-moving number density of a factor of 5.8 at z=1 (for
an assumed q0=0.15 value of the cosmological deceleration
parameter).
(2) Active Galactic Nuclei, which are described by a
model based again on the Rush et al. (1993) local lumi-
nosity function and assuming pure luminosity evolution:
N (L, z) = N (L0, z), where L(z) = L0 exp(kτ [z]) with
k = 3.
Note that the same model with the additional contri-
bution of a population of high-redshift starbursts (forming
elliptical and S0 galaxies as described by Franceschini et
al. 1994) accounts nicely for the cosmological far-IR back-
ground recently detected in the far-infrared and submil-
limeter wavebands by Puget et al. (1996), Hauser et al.
(1997b) and Fixsen et al (1998).
It is also interesting to compare the 12µm counts de-
scribed here with the 15µm counts from the HDF. These
counts are derived from two different ISOCAM filters
(LW10 and LW3) with rather different response functions.
As can be seen in Fig. 6, there appears to be a clear offset
between the two differential galaxy counts by roughly a
factor 2 -- 4 (though the two bins with overlapping fluxes
are in formal agreement). No simple model can explain
this shift in the counts by such a large factor over such
a narrow flux interval. We interpret this shift as proba-
bly not due to actual changes in the counts, but to the
different responses of the LW10 and LW3 filters to the
complex SEDs of galaxies in the mid-IR. Specifically the
7µm PAH emission feature, which enters the LW3 15µm
band at z∼0.5 to 1, and the 10µm absorption feature.
Unfortunately, the strength of these mid-IR spectral fea-
tures varies considerably from object to object (see eg.
Elbaz et al., 1998). A full understanding of the effects
of these features on mid-IR number counts thus awaits
a better theoretical treatment, a better understanding of
the variation of these features locally, and a better idea of
the nature of objects making up the milli -- Jansky 12µm
source population. Xu et al. (1998) used a three compo-
nent model including cirrus, starburst and AGN contri-
butions to fit the mid-IR SEDs of a large sample of local
galaxies. They then extrapolate from this to predict the
effects of the mid-IR SEDs on number counts under vari-
ous evolutionary assumptions. Such an approach may be
useful for understanding the present work and its relation
to the ISOHDF data. However, assumptions would have to
be made about the nature of the faint mid-IR galaxy pop-
ulation, and whether it was significantly different from the
local galaxies studied in detail by IRAS and ISO. There
are already suggestions from the ISOHDF that there are
more and bigger starbursts in the faint mid-IR selected
galaxies than in the local population (Rowan-Robinson et
al., 1997). At this stage we lack redshifts, and thus lu-
minosity and star-formation-rate estimates, for our 12µm
galaxies. A large number of assumptions would thus need
to be made about these objects for an empirical approach
Fig. 5. 12µm Integral Counts for Stars.
Same as in Fig. 4, but for the 13 sources identified as stars in
the 12µm survey. The fluxes are corrected by a factor 1/1.25
to account for the wide LW10 response function and for the
typical SEDs of stars. A comparison is made with the model
stellar counts from Franceschini et al. (1991).
dN (S12)/dS12 versus S12. This is done in Fig. 6, where we
report the Euclidean-normalised differential counts from
our 12µm survey compared to the 15µm number counts
derived from ISOCAM observations of the Hubble Deep
Field (Paper I). The lines correspond to predictions for
Euclidean-normalised counts based on both non-evolving
and strongly evolving population models.
The no-evolution model is based on the local lumi-
nosity function (Saunders et al., 1990) at 60µm and on
the Rush et al (1993) results at 12µm. (for more details
see Franceschini et al. 1997). This minimal curve signifi-
cantly under predicts the observed counts from ISO. We
thus appear to have detected evolution in the 12µm source
population at a 3.5σ significance level.
On the other hand, our observed 12µm counts are
matched by a model assuming an evolving luminosity
function. This is described in terms of two populations,
which we assume dominate the extragalactic sky at these
wavelengths:
(1) Gas-rich systems, i.e. spiral, irregular and star-
bursting galaxies, with luminosity functions evolving with
cosmic time as N (L, z) = N (L, z = 0) exp(k τ (z)),
where N (L, z = 0) is the locally observed distribution (see
above), τ (z) is the lookback time (t0 − t)/t0, where t is the
age of the universe at a redshift z and t0 is the present age,
and k = 3 is the evolution parameter. This corresponds to
density evolution, yielding an average increase in galaxy
Deep 12 Micron Survey with ISO
9
Fig. 6. Euclidean-normalised Differential Number Counts for Galaxies from the 12µm Survey
The differential counts of sources identified with galaxies in the LW10 survey are normalised here to the Euclidean law S−2.5.
The filled squares are from our 12 µm survey, the open circles from the 15 µm survey in the Hubble Deep Field (all with
Poissonian error bars). The number associated with each bin indicates the number of sources in the bin. The dashed line is the
no-evolution curve, while the continuous line corresponds to the model with evolution. The dotted line provides a comparison
with the expected stellar counts (see the corresponding integral counts in Fig. 4).
similar to Xu et al. (1998) to be applied. There would thus
be considerable uncertainties in such an analysis.
A proper test of models of this population is thus even
more critically dependent on obtaining the redshifts of in-
dividual sources than similar work at optical or far-IR
wavelengths. We have therefore begun a followup pro-
gramme to identify and determine the redshifts for all the
12µm galaxies discussed in this paper. Once this data is
available, we will be able to draw firmer conclusions about
10
D.L. Clements et al.
the nature and evolution of the mJy 12µ source popula-
tion.
Our number counts can directly set a lower limit to
the extragalactic infrared background light between 8 and
15 µm. By integrating the light from the galaxies with
a flux larger than 0.5 mJy, we find that νIν (EBL12µm) >
0.50±0.15 nWm−2sr−1. An upper limit of 468nWm−2sr−1
has been reported by Hauser et al. (1998) from DIRBE
(COBE) measurements, which are hampered by the zodi-
acal light.
5. Conclusions
We have performed a deep survey at 12µm using the CAM
instrument on the ISO satellite. We have detected 50 ob-
jects to a 5σ flux threshold of ∼ 500µJy. in a 0.1 deg2
area, of which 13 appear to be stars on the basis of opti-
cal images and optical-IR colours. The remaining 37 ob-
jects appear to be galaxies. We have examined the source
count statistics for this population and find evidence for
evolution in this population, while for stars the counts
are consistent with current galactic structure models and
extrapolations of the optical luminosity functions to the
mid-IR.
Our galaxy counts, when compared with the deep ISO-
CAM counts in the Hubble Deep Field using the LW3
15µm filter, also show evidence for significant effects from
the complex mid-infrared features in the spectral energy
distributions of galaxies.
Acknowledgements. It is a pleasure to thank Herve Aussel,
David Elbaz, Matt Malkan and Jean-Loup Puget for helpful
comments and contributions. The Digitised Sky Survey was
produced at the STSCI under US Government Grant NAG W --
2166. This research has made use of the NASA/IPAC Extra-
galactic Database (NED) which is operated by the Jet Propul-
sion Laboratory, California Institute of Technology, under con-
tract with the National Aeronautics and Space Administration.
The USNO-A survey was of considerable help, and we would
like to express our thanks to all those who helped to produce
it. DLC and ACB are supported by an ESO Fellowship and by
the EC TMR Network programme, FMRX-CT96-0068.
References
Cowie, L.L., Gardner, J.P., Hu, E.M., Songaila, A., Hodapp,
K-W., Wainscoat, R.J., 1994, ApJ, 434, 114
Davies et al., 1997, Icarus, 127, 251
Desert, F-X., Puget, J-L., Clements, D.L., Perault, M.,
Abergel, A., Bernard, J-P., Cesarsky, C.J., 1998, A&A, in
press, Paper I
Elbaz et al., 1998, to appear in Proceedings of the Liege NGST
Workshop, astro-ph/9807209
Fang et al., 1996, BAAS, 189, 5101
Fixsen, D.J., Dwek, E., Mather, J.C., Bennett, C.L., Shafer,
R.A., ApJ., in press
Franceschini, A., de Zotti, G., Toffolatti, L., Mazzei, P.,
Danese, L., 1991, A&AS, 98, 285
Franceschini, A., Mazzei, P., de Zotti, G., Danese, L., ApJ.,
427, 140
Franceschini, A., p. 509, in Extragalactic Astronomy in the In-
frared, proceedings of the 17th Moriond Conference, ed.
Mamon, G.A., Thuan, T.X., Van, J.T.T., pub. Edition
Frontieres, Paris
Giavalisco, M., Steidel, C.C., Macchetto, F., 1996, ApJ., 470,
189
Gregorich, D.T., Neugebauer, G., Soifer, B.T., Gunn, J.E.,
Herter, T.L., 1995, AJ, 110, 259
Hacking, P., 1987, PhD Thesis
Hacking, P., Houck, J. R., 1987, ApJSS, 63, 311
Hauser, M. G., Kelsall, T., Leisawitz, D., & Weiland, J. L.,
1997a, eds., COBE Diffuse Infrared Background Experi-
ment Explanatory Supplement, version 2.1, COBE refer-
ence publication #97-A (Greenbelt: NASA/GSFC), avail-
able in electronic form from the National Space Science
Data Center
Hauser et al., 1997b, BAAS, 191, 910
Hauser et al., 1998, ApJ, in press, Astro-ph//9806167
Joseph, R.D., Wright, G.S., 1985, MNRAS, 214, 87
Kawara, K., et al., 1998, A&A, submitted.
Kessler, M.F., Steinz, J.A., Anderegg, M.E., et al. 1996, A&A,
315, L27
Klemola, A.R., Jones, B.F., Hanson, R.B., 1987, AJ, 94, 501
Lawrence, A., Rowan-Robinson, M., Leech, K., Jones, D.H.P.,
Wall, J.V., 1989, MNRAS, 240, 329
Lilly, S.J., Le Fevre, O., Hammer, F., Crampton, D., 1996,
ApJ., 460, L1
Madau, P., Pozzetti, L., Dickinson, M., 1998, ApJ., 498, 106
Monet, D., et al., 1996, USNO-A Catalogue, US Naval Obser-
vatory, Flagstaff Station
Murdoch, H.S., Crawford, D.F., Jauncy, D.L., 1973, ApJ., 183,
1
Abraham, R.G., Tanvir, N.R., Santiago, B.X., Ellis, R.S.,
Glazebrook, K., van den Bergh, S., 1996, MNRAS, 279,
L47
Aussel, H., Cesarsky, C. J., Elbaz, D. and Starck, J.-L., 1998,
A&A, submitted
Bertin, E., Dennefeld, M., Moshir, M., 1997, A&A, 323, 685
Cesearsky, C.J., et al., 1994, The ISOCAM Observers Manual,
ESA, Noordwijk.
Cesarsky, C.J., et al., 1996, A&A, 315, L32
Clements, D.L., Sutherland, W.J., Saunders, W., McMahon,
R.G., Maddox, S., Efstathiou, G.P., Rowan-Robinson, M.,
1996a, MNRAS, 279, 459
Clements, D.L., Sutherland, W.J., Saunders, W., McMahon,
R.G., 1996b, MNRAS, 279, 477
Okuda, H., in Proceedings of Astrophysics with Infrared Sur-
veys, 1998, ed. Bicay, M., in press
Oliver, S.J., 1995, in Wide-Field Spectoscopy and the Distant
Universe, eds. Maddox, S.J., & Aragon-Salamanca, pub.
World Scientific
Oliver, S.J., et al., 1997, MNRAS, 289, 471
Puget, J-L., Abergerl, A., Bernard, J-P., Boulanger, F., Bur-
ton, W.B., Desert, F-X., Hartmann, D., 1996, A&A, 308,
L5
Puget, J-L., et al, 1999, A&A, in press
Rowan-Robinson, M., Broadhurst, T., Oliver, S.J., Taylor,
A.N., Lawrence, A., McMahon, R.G., Lonsdale, C.J., Hack-
ing, P.B., Conrow, T., 1991, Nat., 352, 677
Rowan-Robinson et al., 1997, MNRAS, 289, 490
Deep 12 Micron Survey with ISO
11
Rush, B., Malkan, M., Spinoglio, L., 1993, ApJS., 89, 1
Sanders, D.B., & Mirabel, I.F., ARAA, 34, 749
Saunders, W., Rowan-Robinson, M., Lawrence, A., Efstathiou,
G., Kaiser, N., Ellis, R.S., Frenk, C.S., 1990, MNRAS, 242,
318
Serjeant, S., Lacy, M., Rawlings, S., King, L.J., Clements, D.L.,
1995, MNRAS, 276, L31
Steidel, CC., Giavalisco, M., Pettini, M., Dickinson, M., Adel-
berger, KL., 1996, ApJ., 462, L17
Stockman, P., Mather, J., 1997, BAAS, 191, 5403
Sykes, M.V., Walker, R.G., 1992, Icarus, 95, 180
Werner, M.W., Bicay, M.D., 1997, BAAS, 191, 4207
Xu, C., Hacking, P.B., Fang, F., Shupe, D.L., Lonsdale, C.J.,
Lu, N.Y., Helou, G., ApJ., 508, 576
12
D.L. Clements et al.
F12(µJy) RA(2000) DEC(2000)
Class
-9 31 21.6 Extended; IRAS Source
Star
Star
Star
GSC-galaxy; Extended
Star
Star
Star
Binary Star
Star
Star
Binary Star
Star
Star
Star
Name
F1 0
F1 1
F1 2
F1 3
F1 4
F1 5
F1 9
F1 10
F1 11
F1 12
F1 18
F1 22
F1 30
F1 44
F1 48
F2 0
F2 1
F2 3
F2 6
F2 12
F2 16
F2 23
F2 24
F2 80
F2 87
F2 195
F3 0
F3 1
F3 2
F3 3
F3 4
F3 5
F3 9
F3 15
F3 34
F3 37
F4 0
F4 1
F4 2
F4 3
F4 4
F4 5
F4 6
F4 9
F4 11
F4 12
F4 24
F4 28
F4 57
F4 68
12147.± 200.
4258.± 219.
4155.± 131.
3835.± 99.
2150.± 185.
1046.± 102.
3732.± 142.
1903.± 185.
1324.± 113.
842.± 92.
2190.± 171.
615.± 112.
636.± 88.
622.± 109.
605.± 119.
10479.± 260.
6884.± 175.
2385.± 100.
911.± 99.
744.± 136.
770.± 104.
702.± 95.
814.± 118.
522.± 89.
536.± 88.
464.± 87.
7265.± 104.
3266.± 98.
1901.± 100.
1760.± 95.
1682.± 105.
1374.± 108.
5444.± 93.
597.± 100.
639.± 105.
648.± 101.
7684.± 140.
3670.± 149.
3581.± 155.
3207.± 84.
3166.± 357.
2481.± 215.
3051.± 197.
1234.± 95.
806.± 89.
933.± 106.
559.± 92.
474.± 88.
615.± 118.
690.± 87.
3 5 36.2
3 5 4.5
3 5 40.4
3 5 15.1
3 5 5.3
3 5 24.5
3 5 35.9
3 5 6.1
3 5 14.9
3 5 28.4
3 5 36.3
3 5 33.0
3 5 30.7
3 5 36.7
3 5 11.0
3 1 6.1
3 0 36.7
3 0 37.3
3 0 31.0
3 0 59.4
3 0 38.2
3 0 39.7
3 0 35.4
3 0 39.5
3 0 37.6
3 0 35.8
3 9 42.6
3 9 56.7
3 9 52.5
3 10 0.9
3 10 9.1
3 9 44.7
3 9 42.2
3 10 3.3
3 9 46.1
3 9 57.4
3 3 54.5
3 4 3.4
3 3 35.8
3 3 55.2
3 3 53.8
3 3 47.3
3 4 6.3
3 3 56.8
3 4 1.4
3 3 54.7
3 3 58.0
3 3 55.6
3 3 55.8
3 3 55.7
-9 31 6.6
-9 33 56.1
-9 31 17.6
-9 32 7.7
-9 35 50.7
-9 31 43.8
-9 32 41.5
-9 31 3.8
-9 35 17.5
-9 31 32.2
-9 31 50.4
-9 33 13.4
-9 32 30.1
-9 33 10.0
-10 44 27.5
-10 36 57.3
-10 42 51.0
-10 41 7.2
-10 44 38.2
-10 42 46.6
-10 40 47.0
-10 43 22.3
-10 39 43.4
-10 41 59.4
-10 41 6.5
-8 35 33.4
-8 36 24.5
-8 37 1.5
-8 39 39.2
-8 38 57.2
-8 32 55.6
-8 35 50.3
-8 37 46.7
-8 32 44.0
-8 36 44.2
-9 59 6.7
-10 1 18.0
-9 55 43.1
-9 56 59.9
-9 50 40.0
-9 59 18.2
-9 53 47.6
-9 57 31.3
-9 57 36.6
-9 58 45.6
-9 59 5.2
-9 56 11.0
-9 59 28.3
-9 56 40.4
Table 2. 12µm Sources Detected at > 5σ
Objects are classified as stars based on their optical/IR colours (see section 3.1). The object indicated as GSC-galaxy is in
the GSC but inspection of the DSS image reveals it to be a galaxy. Two of the stars F3 0 and F3 9 are very close, precluding
accurate photometry.
Name F12(µJy) Bmag Rmag F12P RED (µJy) Pred/Obs Excess?
Yes
Deep 12 Micron Survey with ISO
13
1794
4962
442
8018
747
540
1220
2289
720
0.42
1.19
0.72
1.16
0.97
0.77
0.64
0.70
1.04
F1 1
F1 2
F1 22
F2 1
F2 16
F2 23
F3 2
F3 1
F4 68
4258
4155
615
6884
770
702
1902
3266
690
12.6
10.9
13.1
13.1
13.7
13.7
13.5
13.5
13.7
11.7
10.3
12.7
11.3
12.7
12.9
12.3
12.0
12.9
Table 3. Predicted and Actual Stellar Fluxes in the 12µm ISO Survey
Predicted fluxes are calculated as described in the text. A star is described as having a candidate IR excess if its predicted flux
is less than half the measured flux.
|
astro-ph/9808007 | 1 | 9808 | 1998-08-03T19:48:03 | Synchrotron Self Absorption in GRB Afterglow | [
"astro-ph"
] | GRB afterglow is reasonably described by synchrotron emission from relativistic blast waves at cosmological distances. We perform detailed calculations taking into account the effect of synchrotron self absorption. We consider emission from the whole region behind the shock front, and use the Blandford McKee self similar solution to describe the fluid behind the shock. We calculate the spectra and the observed image of a GRB afterglow near the self absorption frequency $\nu_a$ and derive an accurate expression for $\nu_a$. We show that the image is rather homogeneous for $\nu<\nu_a$, as opposed to the bright ring at the outer edge and dim center, which appear at higher frequencies. We compare the spectra we obtain to radio observations of GRB970508. We combine the calculations of the spectra near the self absorption frequency with other parts of the spectra and obtain revised estimates for the physical parameters of the burst: $E_{52}=0.53$, $\epsilon_e=0.57$, $\epsilon_B=0.0082$, $n_1=5.3$. These estimates are different by up to two orders of magnitude than the estimates based on an approximate spectrum. | astro-ph | astro-ph |
Synchrotron Self Absorption in GRB
Afterglow
Jonathan Granot, Tsvi Piran and Re’em Sari
Racah Institute, Hebrew University, Jerusalem 91904, Israel
February 1, 2008
Abstract
GRB afterglow is reasonably described by synchrotron emission
from relativistic blast waves at cosmological distances. We perform
detailed calculations taking into account the effect of synchrotron self
absorption. We consider emission from the whole region behind the
shock front, and use the Blandford McKee self similar solution to
describe the fluid behind the shock. We calculate the spectra and the
observed image of a GRB afterglow near the self absorption frequency
νa and derive an accurate expression for νa . We show that the image
is rather homogeneous for ν < νa , as opposed to the bright ring at the
outer edge and dim center, which appear at higher frequencies. We
compare the spectra we obtain to radio observations of GRB970508.
We combine the calculations of the spectra near the self absorption
frequency with other parts of the spectra and obtain revised estimates
for the physical parameters of the burst: E52 = 0.53, ǫe = 0.57, ǫB =
0.0082, n1 = 5.3. These estimates are different by up to two orders of
magnitude than the estimates based on an approximate spectrum.
1 Introduction
The detection of delayed x-ray, optical and radio emission following a GRB,
known as GRB afterglow, is described reasonably well by emission from
a spherical relativistic shell, decelerating upon collision with an ambient
1
medium (Waxman 1997a, Wijers, Rees & M´esz´aros 1997, Katz & Piran 1997,
Sari, Piran & Narayan 1998). A relativistic blast wave expands through the
ambient medium, continuously heating up fresh matter as it passes through
the shock. In these models, the GRB afterglow is the result of synchrotron
emission of the relativistic electrons of the heated matter.
Several recent works considered emission from various regions on or be-
hind the shock front (Waxman 1997c, Sari 1998, Panaitescu & M´esz´aros
1998, Granot, Piran & Sari 1998 (GPS hereafter), Gruzinov & Waxman
1998). These authors considered the spectra near the peak frequency νm and
found that exact calculations of the spectrum could differ by up to one order
of magnitude from simpler estimates.
We consider emission from an adiabatic highly relativistic blast wave ex-
panding into a cold and uniform medium. We consider the effect of the whole
volume behind the shock front, the importance of which was stressed in GPS.
The hydrodynamics is described by the Blandford McKee (1976 denoted BM
hereafter) self similar solution. We consider synchrotron emission and we ig-
nore Compton scattering and electron cooling. Similar to GPS, we consider
several models for the evolution of the magnetic field.
Synchrotron self absorption becomes significant below a critical frequency
νa called the self absorption frequency. We assume that νa ≪ νm , where νm
is the peak frequency, which is reasonable for the first few months after the
burst. While the spectrum near νm was quite extensively studied, so far
only order of magnitude estimates of the spectrum near the self absorption
frequency νa were done.
For a system in which νa ≪ νm , the spectrum, Fν , is proportional to ν 2 for
ν ≪ νa , rather than the standard ν 5/2 , because almost all the low frequency
radiation is emitted by electrons with a typical synchrotron frequency much
higher than νa (Katz 1994). The spectra far above νa is proportional to
ν 1/3 .
In this paper we explore the spectra near νa , and find how the two
asymptotic forms join together. An analysis of the spectra over a wider range
of frequencies was made by Sari, Piran & Narayan (1998), and a detailed
analysis of the spectra near the peak frequency, taking a full account of the
BM solution, was made in GPS and Gruzinov & Waxman (1998).
The physical model is described in §2.
In §3 we describe the compu-
tational formalism. The spectra for several magnetic field models and the
observed image of a GRB afterglow at various frequencies are presented in
§4.
In §5 we compare the calculated spectra to radio observations of the
2
afterglow of GRB970508. When we use the modified calculations of the self
absorption spectra and of the spectra around the peak (GPS) and the cooling
frequency (Sari, Piran & Narayan 1998) we find new estimates of the param-
eters of GRB970508. These estimates are different by more than an order
of magnitude than estimates based on a simpler broken power law spectra
(Wijers & Galama 1998).
2 The Physical Model
The underlying model assumes an ultra-relativistic spherical blast wave ex-
panding into a cold and uniform medium. The blast wave constantly heats
fresh matter, and the observed afterglow is the result of synchrotron emission
of the relativistic electrons of the heated matter. We consider an adiabatic
evolution, where the fluid behind the shock is described by the BM self
similar solution. It has been numerically verified that for an adiabatic evolu-
tion with general initial conditions the solution approaches the BM solution
(Kobayashi, Piran & Sari 1998). The BM solution is expressed in terms of
the similarity variable χ, which is defined by:
f (cid:18) R − r
R (cid:19) ,
χ ≡ 1 + 16γ 2
where R is the radius of the shock front, r is the distance of a point from the
center of the burst and γf is the Lorentz factor of the matter just behind the
shock. The BM solution is given by:
(1)
n′ = 4γf n0χ−5/4
,
γ = γf χ−1/2
,
e′ = 4n0mp c2γ 2
f χ−17/12 ,
(2)
where n′ and e′ are the number density and the energy density in the local
frame, respectively, γ is the Lorentz factor of the bulk motion of the matter
behind the shock, mp is the mass of a proton and n0 = n1 × 1cm−3 is the
proper number density of the unshocked ambient medium.
We consider three alternative models for the magnetic field: B , B⊥ and
B ′ = ǫB e′ (i.e. equipartition) everywhere. For the two
Brad . B satisfies e′
B ′ = ǫB e′ on the shock
other magnetic field models we assume they acquire e′
front, and from then on, evolve according to the “frozen field” approximation.
This implies that the magnetic fields equal Bχ3δ/2 , where δ = 7/18 when the
3
,
(3)
ν ′ ,e ∼=
P ′
K = (p − 1)n′γ p−1
min .
magnetic field is in the radial direction (Brad ) and δ = −7/36 when the
magnetic field is perpendicular to the radial direction (B⊥ ) (see GPS).
We assume that the energy of the electrons is everywhere a constant
el = ǫee′ , and that the shock produces a
fraction of the internal energy: e′
−p 1 for γe ≥ γmin . We use
power law electron distribution: N (γe ) = K γe
p = 2.5 wherever a definite numerical value of p is needed. The constants
K and γmin in the electron distribution can be calculated from the number
density and energy density:
p − 1 ! ǫee′
γmin = p − 2
n′me c2
We consider frequencies which are much lower than the typical syn-
chrotron frequency ν ≪ νsyn , where the spectral power of an electron emitting
synchrotron radiation can be approximated by:
syn !1/3
ν ′
ν ′
(Rybicki & Lightman 1979) where Γ is the gamma function, B ′ is the mag-
netic field, me and qe are the mass and the electric charge of the electron,
respectively, and α is the angle between the directions of the electron’s ve-
locity and the magnetic field, in the local frame.
Rl and γl are the location of the shock and its Lorentz factor on this
line of sight and Rl /γl is an estimate of the shell thickness in the local frame.
The optical depth along the LOS can be approximated by α′
ν ′ Rl /γl where the
absorption coefficient α′
ν ′ is taken at y = χ = 1. The “back of the envelope”
estimate for the self absorption frequency ν0 is therefore the frequency which
satisfies: α′
Rl /γl = 1. ν0 is given by:
ν ′
0
3p + 2 !3/5 (p − 1)8/5
ν0 = 4.24 × 109(1 + z)−1 p + 2
(p − 2)
where z is the cosmological red shift and E = E52 × 1052ergs is the total
energy of the shell. We also define a “standard” flux density F0 , which is an
3γ 2
e qeB ′ sin α
4πme c
25/3π
Γ(1/3)
q 3
e B ′ sin α
me c2
B E 1/5
e ǫ1/5
52 n3/5
ǫ−1
1 Hz ,
,
ν ′
syn ≡
,
(4)
(5)
1 for the energy of the electrons to remain finite we must have p > 2.
4
Fm ≡
,
(1 + z)
3
approximate expression for the flux density at the self absorption frequency.
The peak flux and peak frequency can be approximated by:
n1R3
L γP ′
syn ! (y = χ = 1)
e
l
νm ≡ νsyn (γmin , y = χ = 1) ,
d2
ν ′
(6)
e = (4/3)σT cγ 2
(see F0 and νT in GPS) where P ′
e e′
B is the total emitted power
of an extreme relativistic electron (Rybicki & Lightman 1979), σT is the
Thomsom cross section and dL = dL28 × 1028cm is the luminosity distance 2 .
Now we can define: F0 ≡ Fm(νa/νm )1/3 :
3p + 2 !1/5 (p − 1)6/5
F0 = 1.31(1 + z) p + 2
(p − 2)
where Tdays is the observed time in days.
We express the observed frequency ν in units of ν0 : ν ≡ φν0 , thus intro-
ducing the dimensionless variable φ which we use to express our results.
1 T 1/2
52 n7/10
B E 9/10
e ǫ2/5
L28 ǫ−1
d−2
daysmJy , (7)
3 The Formalism
We consider a system that is moving relativistically while emitting radiation,
and obtain a formula for the flux density measured by a distant observer,
allowing for self absorption. We denote quantities measured in the local
rest frame of the matter with a prime, while quantities without a prime are
measured in the observer frame.
The specific intensity Iν (energy per unit time per unit area per unit
frequency per unit solid angle) satisfies the radiative transfer equation:
dIν
ds
= jν − αν Iν ,
where jν and αν are the emission coefficient and the absorption coefficient,
respectively, and s is the distance along the beam. When self absorption
becomes important the contribution to Iν depends on the optical depth along
the path within the emitting system and Iν should be integrated separately
for every tra jectory.
(8)
2Here and latter we use a general cosmological model. For a given model we should
substitute the appropriate expression for dL (1 + z , H0).
5
A
r
θ
P
R⊥
B
Rl
Figure 1: The egg-Shaped curve is the boundary of the region from which
photons reach a distant observer simultaneously. The observer is located far
to the right and the symmetry axis is the LOS from the source of the burst to
the observer. The bold line represents the tra jectory of a photon that reaches
the observer. For a distant observer these tra jectories are almost parallel to
the LOS, and are therefore characterized by their distance R⊥ from the LOS.
Iν = 0 at point A, and reaches its final value at point B. A photon emitted
at a point P can be absorbed or cause stimulated emission at any point along
the tra jectory, until it passes the shock front at point B.
For the BM solution, the radius of the shock front and the Lorentz factor
of the matter just behind the shock, on the LOS for a given observed time
T , are given by:
T 1/4
days cm .
(9)
T −3/8
days
γl ∼= 3.65E 1/8
, Rl ∼= 5.53 × 1017E 1/4
52 n−1/8
52 n−1/4
1
1
For a given observed time T , we use the coordinates y and χ, instead of the
polar coordinates r and θ, where y is defined by y ≡ R/Rl . R is the radius of
the shock front at the time t at which a photon must be emitted in order to
reach the observer at the observed time T . A photon emitted at a location
(r, θ) at a time t in the observer frame, reaches the observer at an observed
time T given by:
rµ
T = t −
c
where µ ≡ cos θ (for more details on the coordinates, see GPS).
Since the observer is distant, the tra jectory of a photon that reaches the
observer is almost parallel to the LOS in the observer frame (see Figure 1).
,
(10)
6
T 5/8
days cm ,
(11)
(12)
(13)
R⊥
R⊥,max
=
We can therefore parameterize the various tra jectories by their distance from
the LOS. The distance R⊥ of a point from the LOS is given by:
√2Rl
4γl qy − χy 5 .
R⊥ ≡ r sin θ ∼= Rl yq1 − µ2 ∼=
The maximal value of R⊥ is obtained for χ = 1, y = 5−1/4 and is given by:
R⊥,max = 3.91 × 1016E 1/8
52 n−1/8
1
We express R⊥ in units of R⊥,max , introducing the dimensionless variable:
55/8
2 qy − χy 5 .
x ≡
In order to solve equation 8, explicit expressions for the absorption and the
emission coefficients are needed. We consider an isotropic electron velocity
ν ′ ,e ∝ sin2/3 α we use the averaged value hsin2/3 αi =
distribution, and since P ′
√πΓ(1/3)/5Γ(5/6). The total power per unit volume per unit frequency in
the local frame is given by:
ǫ1/3
B γln4/3
(p − 1)5/3
0 q 8/3
e ν ′1/3
ν ′ = Z ∞
P ′
m1/3
(3p − 1)(p − 2)2/3
p cy 3/2χ29/18−δ
γmin
(14)
Assuming the emission and absorption are isotropic in the local rest frame
ν ′ = P ′
of the matter, j ′
ν ′ /4π and the absorption coefficient is given by:
(p + 2)
N (γe )
8πmeν ′2 Z ∞
γe
γmin
ǫ1/3
b n4/3
(p + 2)(p − 1)8/3
0 q 8/3
e
ǫ5/3
e m4/3
(3p − 2)(p − 2)5/3
p cν ′5/3χ13/9−δ
(Rybicki & Lightman 1979). Keeping in mind that jν /ν 2 and αν ν are Lorentz
invariant and ν ′ = ν γ (1 − β µ), we can write the radiative transfer equation
(equation 8) explicitly. We first write an expression for the optical depth to
the observer, which will later help us gain some intuition for the results. The
optical depth to the observer is given by:
2
φ5/3 Z ymax (x)
τ (x, y ) = Z αν ds =
y
y 5/3
χ10/9−δ (x, y )(1 + 7χ(x, y )y 4)2/3 ,
64π 13/632/3
5Γ(5/6)
dγeP ′
ν ′ ,e(γe )
=
8π 7/632/3
5Γ(5/6)
dγeN (γe )P ′
ν ′ ,e =
.
(15)
(16)
α′
ν ′ =
,
dy
7
where χ(x, y ) is obtained from equation 13, and ymax (x) is obtained by solving
χ(x, y ) = 1 for y .
We define a “typical” specific intensity I0 by: I0 ≡ Sν (y = χ = 1), where
the source function Sν is defined as: Sν ≡ jν /αν . Expressing Iν in units of
I0 : Iν ≡ Iν I0 , we write equation 8 in terms of the dimensionless variables Iν :
d Iν
2y 5/3
φ5/3χ10/9−δ (1 − 7χy 4)2/3 8yχ1/3
(1 − 7χy 4) − Iν ! ,
dy
where we took ds ∼= Rldy . This equation can be solved numerically for a
given magnetic field model and a given value of φ, for different values of x,
thus obtaining Iν (x). The observed flux density is given by:
!2
Z dS⊥Iν = 2π(1 + z) R⊥,max (T )
Z 1
dL
0
As can be seen from equation 18, the surface brightness is proportional to
Iν : dFν /dS⊥ ∝ Iν . This means that we also obtain the observed image of a
GRB afterglow, by calculating the observed flux density in this way.
xdxIν (x, T ) . (18)
(17)
=
Fν (T ) ∼=
(1 + z)
d2
L
4 The Spectra and Observed Image
Solving equations 17 and 18, we calculate the spectra for the three mag-
netic field models (B , Brad and B⊥ ). These spectra are shown in Figure
2. We define the self absorption frequency νa as the frequency at which the
asymptotic high and low frequency power laws meet. For the equipartition
magnetic field model B we obtain:
B E 1/5
e ǫ1/5
52 n3/5
ǫ−1
1 Hz ,
3p + 2 !3/5 (p − 1)8/5
νa = 0.247ν0 = 1.05 × 109(1 + z)−1 p + 2
p − 2
(19)
while for the B⊥ model it is lower by 6% and for the Brad model it is higher
by 14%. For ν ≪ νa the system is optically thick , and therefore the ra-
diation reaching an observer is essentially emitted near the edge facing the
observer. It reflects the electron distribution there, and it is independent of
the magnetic field model. For ν > νa the system is optically thin and the
flux density is different for the various magnetic field models.
8
100
10−1
0
F
/
ν
F
10−2
10−3
Fν
,ext(B)
a
Brad
B
B⊥
νa(B)
10−1
100
φ
101
Figure 2: The spectra for different magnetic field models. The frequency νa
and the flux density Fνa ,ext are defined at the point where the extrapolations
of the power laws meet, as is illustrated for the equipartition magnetic field
model B . νa is constant in time, Fνa ,ext ∝ T 1/2 and both don’t change
significantly between the different magnetic field models. For φ ≪ 1 (ν ≪
νa ) the system is optically thick, and the flux density reflects the electron
distribution (or the Lorentz boosted “effective temperature” of the electrons)
and is independent of the magnetic field model.
9
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 3: The optical depth τ to the observer, divided by the maximal
optical depth (τ = τmax = 1.08 × (ν /νa )−5/3 is black). τmax is obtained at
x ≡ R⊥/R⊥,max = 0.93. The contour lines are equally spaced with a 5%
interval between following contour lines.
The ratio τ /τmax (where τ is the optical depth and τmax is its maximal
value) for the equipartition B model is shown in Figure 3. Since τν ∝ φ−5/3
everywhere, τ /τmax is frequency independent. The contour lines of τν are
dense where the absorption coefficient αν is large. τmax is obtained at x ∼=
0.93, i.e. quite close to the edge of the image, since then the whole tra jectory
to the observer is relatively close to the shock front, implying large values of
αν and a large contribution to τν . τmax is a good indicator for the opacity
of the system, and for the equipartition B model it is given by: τmax =
1.08 × (ν /νa )−5/3 .
We define Fνa ,ext as the extrapolated flux density at νa (see Figure 2).
For the equipartition B model we obtain:
3p + 2 !1/5 (p − 1)6/5
Fνa ,ext = 0.108F0 = 142(1+z) p + 2
(p − 2)
(20)
While for the B⊥ model it is lower by 12% and for the Brad model it is
higher by 23%. The actual flux density at νa is around 35% lower than
1 T 1/2
52 n7/10
B E 9/10
e ǫ2/5
d−2
L28ǫ−1
daysµJy ,
10
Fνa ,ext: Fνa ∼= 0.65Fνa ,ext . The values of Fνa ,ext and νa are useful, since for
ν ≪ νa magnitude smaller than νa (assuming νm > νa ) the flux density
is given by: Fν ∼= Fνa ,ext(ν /νa )2 , and for νa ≪ ν ≪ νm it is given by:
Fν ∼= Fνa ,ext(ν /νa )1/3 . Both approximations are already good within a few
percent for frequencies a factor ∼ 3 below or above νa , respectively.
It is interesting to compare the spectra obtained for the BM solution
to that obtained for a simplistic model of a static homogeneous disk. This
comparison should help us learn whether the fact that the spectra is rounded
up near νa should be attributed mainly to the specific hydrodynamics used,
or whether it is a more general feature of a calculation accounting for self
absorption. For a static disk we obtain the well known result:
νa ,extψ 2 (cid:16)1 − exp[−ψ−5/3 ](cid:17)
Fν = F ∗
where the constants F ∗
νa ,ext and ν ∗
a are determined by the radius and width
of the disk and by the hydrodynamic parameters of the emitting matter
within the disk. If these parameters are set so that the two stared quantities
equal those in equations 20 and 19, respectively, the resulting flux density
is very similar to that obtained for the BM solution. Substituting equations
20 and 19 into equation 21 can thus serve as a good approximation (better
than 3%) for the observed flux density at ν ≪ νm . This similarity implies
that the shape of the spectrum near νa is rather independent of the specific
hydrodynamic solution considered. On the other hand, the exact values of
νa and Fνa ,ext depend significantly on the hydrodynamics, and can not be
determined without the detailed calculation. A simplified calculation for ν ∗
a
could for example yield ν0 (equation 5) instead of νa that is given in equation
19.
, ψ ≡
(21)
ν
ν ∗
a
,
The surface brightness as a function of x ≡ R⊥/R⊥,max for the equipar-
tition magnetic field model B is shown in Figure 4, for a few representative
values of φ ≡ ν /ν0 . An illustration of the observed image of a GRB after-
glow, which is implied from the surface brightness, is shown in Figure 5. As a
reference, the image for ν ≫ νm , which was derived in GPS, is also presented.
For ν ≫ νa the system is optically thin and the result coincides with that
obtained in GPS for ν ≪ νm . There is a bright ring near the outer edge of
the image, and the surface brightness at the center of the image is 58% of its
average value. As ν decreases, the contrast between the edge and the center
of the image decrees and the system becomes increasingly optically thick.
11
1.8
1.6
1
s
1.4
s
e
n
1.2
t
h
g
i
r
b
0.8
e
c
0.6
a
f
r
u
0.4
s
0.2
0
0
ν > ν
a
ν < ν
a
0.2
0.4
0.6
x ≡ R⊥ / R⊥,max
0.8
1
Figure 4: The surface brightness divided by the average surface brightness,
as a function of x, for Log10(φ) = −1.5, −1 − 0.75, −0.5, 0 (corresponding to
ν /νa = 0.13, 0.40, 0.72, 0.28, 4.05). At high frequencies the contrast between
the center and the edge of the image is larger than at low frequencies. At low
frequencies the system is optically thick, and the surface brightness reflects
the “effective temperature” of the electrons at the edge of “egg” depicted in
Figure 1, on the side facing the observer.
12
ν/νa = 0.1
ν/νa = 1
ν/νa = 10
ν >> νm
Figure 5: The observed image of a GRB afterglow, at several frequencies.
At high frequencies there is a bright ring near the outer edge of the image,
and the contrast between the center and the edge of the image is larger than
at low frequencies. At low frequencies the surface brightness increases as
one moves from the center towards the edge, until it drops very sharply, due
to the fact that the system becomes optically thin near the edge. The last
image, for ν ≫ νM , is taken from GPS and it is brought as a reference, to
illustrate the change in the relative surface brightness along the image over
a large range of frequencies. At ν ≫ νm there is a thin bright ring at the
outer edge of the image, and the surface brightness at the center is only a
few percent of its maximal value.
13
As a result, the surface brightness reflects the Lorentz boosted “effective
temperature” of the electron distribution, at the outer edge of the “egg”
depicted in Figures 1 and 3, on the side facing the observer. Larger values
of x correspond to smaller shock radii R, and since the electrons posses a
larger Lorentz factor (i.e. a larger “effective temperature”) at smaller radii,
the surface brightness increases with x. This is true as long as long as the
optical depth is still large. Since the length of the tra jectory within the
system approaches zero as x → 1, for every given frequency ν the system
becomes optically thin for values of x sufficiently close to 1. For frequencies
smaller than νa by more than one or two orders of magnitude, this occurs at
1 − x ≪ 1 so that the drop in the surface brightness near x = 1 is extremely
sharp. For ν ≪ νa the surface brightness at the center of the image is 77%
of its average value, resulting in an almost uniform disk, rather than a ring
which is obtained for ν > νa . The uniformity of the image for ν ≪ νa stands
out even more when compared to ν ≫ νm , where there is a thin bright ring
on the outer edge of the image and the surface brightness at the center is
only a few percent of its maximal value.
5 Comparison to Observations
We now fit the theoretical spectra calculated, to the radio frequencies obser-
vations of the afterglow of GRB970508 (Fig. 4 of Shepherd 1998). Perhaps a
future near enough GRB will result in a sufficient resolution, that will enable
us compare the predicted images of a GRB afterglow to observations.
Since both ν0 and F0 depend on the parameters of the model (see equa-
tions 5 and 7), we have two degrees of freedom in trying to fit the calculated
spectra to the observed data. We fit the equipartition magnetic field model
B to the data in Shepherd et al (1998), and obtain:
νa = 3.1 ± 0.4 × 109Hz
, Fνa ,ext = 450 ± 37µJy ,
with χ2/dof = 0.48. The fit is presented in Figure 6.
The errors quoted are statistical errors. The actual errors are probably
larger due to uncertainty in the radio flux due to scintillation, and due to
the fact that not all the observations are exactly simultaneous (see Shepherd
1998).
(22)
14
Fν
,ext
a
103
]
y
J
µ
[
ν
102
F
νa
101
109
1010
ν [Hz]
1011
Figure 6: A fit of the calculated spectra, to radio observations of the afterglow
of GRB970508. We obtained: νa = 3.1 ± 0.4 × 109Hz and Fνa ,ext = 450 ±
37µJy.
Substituting these results in equations 19 and 20, respectively, we obtain
two constraints on the physical parameters of the burst. The luminosity
distance depends on the cosmological model. For Ω = 1, Λ = 0 (which
implies: dL = 2(1 + z − √1 + z )c/H0), H0 = 65Kms−1Mpc−1 and z = 0.835
we find:
3p + 2 !3/5
(p − 1)8/5
(p − 2) p + 2
3p + 2 !1/5
(p − 1)6/5
(p − 2) p + 2
B E 9/10
e ǫ2/5
52 n7/10
ǫ−1
1
Substitution of the value for νa from the data (equation 22) into equation
19, yields an equation different by a factor of ∼ 2 from equation 22 of Wijers
& Galama (1998). The difference is mainly due to a different theoretical
value they used for νa . They also used a slightly different value of νa as
corresponding to the same observational data. This factor of ∼ 2 implies
significant corrections to the values of the physical parameters of the burst
B E 1/5
e ǫ1/5
52 n3/5
ǫ−1
1
= 5.43 ,
(23)
(24)
= 1.22 .
15
TABLE 1
Estimates of the Physical Parameters of GRB970508.
model
E52
ǫe
ǫB
n1
broken
power law
3.7
0.13
0.068
0.035
modified modified νa ,
νm , Fνm , νc
νa
0.53
2
0.57
0.24
0.011
0.0082
5.3
0.70
Table 1: The first row depicts the values of the physical parameters of
GRG970508 as calculated by Wijers & Galama (1998). The other rows show
how these values change when some of the equations they used in the cal-
culation are corrected. The first line lists the measurable quantities, whose
equations were altered.
(see the second row of Table 1). For example, n1 (or n in their notation)
becomes a factor of ∼ 20 larger: n1 = 0.7 instead of n1 = 0.035.
We modify the estimates further using more detailed calculations of the
spectrum near the peak flux (GPS), and use a different estimate for the
cooling frequency νc (Sari, Piran & Narayan 1998). We discover that the
values obtained for the physical parameters of a burst are very sensitive to
the theoretical model of the spectrum. In order to illustrate this we show
in Table 1 the physical parameters of GRB970508 for different estimates of
the spectrum. The first column uses a broken power law spectrum (Wijers
& Galama 1998); the second column uses a corrected theoretical value for νa
from equation 19 and a corrected observational value for νa from equation 22;
the third column adds a modified theoretical values to νm and Fνm (which are
taken from GPS - denoted there as νpeak and Fν,max ) and a different estimate
for the cooling frequency νc (from Sari, Piran & Narayan 1998), keeping the
observational values from Wijers & Galama (1998).
Our best values (with all modifications added) are E52 = 0.53, ǫe = 0.57,
ǫB = 0.0082 and n1 = 5.3. These values differ by one to two orders of
magnitude than the values obtained by Wijers & Galama (1998) using a
16
broken power law spectra. One must keep in mind that it is difficult to
obtain an accurate estimate of the various observables from the observed
data. It is especially difficult to determine νm , Fνm and νc . Therefore, more
than obtaining a better estimate of the physical parameters of the burst,
these calculations show the sensitivity of this method and the need for more
accurate data.
Sufficient observational data has been gathered on the radio afterglow
of GRB980329, to enable a fit similar to the one we made for GRB970508.
Such a fit was carried out by Taylor et al (1998). The theoretical formula for
the flux density that was used for the fit is identical to equation 21. Since
this is a good approximation for the shape of the spectra near νa , we can
use the values extracted from the data to obtain constraints on the physical
parameters of GRB980329.
The values used by Taylor et al for the flux density are mean values over
the first month. Since Fνa ∝ T 1/2 , this corresponds to the flux density at
∼ 15 days. The values extracted from the data are:
, Fνa ,ext ∼= 600µJy .
νa ∼= 1.3 × 1010Hz
Substituting these results in equations 19 and 20, respectively, we obtain
two constraints on the parameters of GRB980329. For Ω = 1, Λ = 0 and
H0 = 65Kms−1Mpc−1 we find:
3p + 2 !3/5
(p − 1)8/5
(p − 2) p + 2
2
(1 + z)
√2 − 1
√1 + z − 1 !2
3p + 2 !1/5
(p − 1)6/5
(p − 2) p + 2
B E 9/10
e ǫ2/5
52 n7/10
ǫ−1
1
where the red shift z of this burst is not yet known.
∼= 26 , (26)
(25)
B E 1/5
e ǫ1/5
52 n3/5
ǫ−1
1
∼= 1.5 , (27)
6 Discussion
We have considered synchrotron emission from a system moving relativisti-
cally, taking into account the effect of synchrotron self absorption. We have
assumed an adiabatic evolution and used the self similar solution of Bland-
ford & McKee (1976) to describe the matter behind the shock. This solution
describes an extreme relativistic spherical blast wave expanding into a cold
17
uniform medium. Our calculations accounted for the emission from the whole
region behind the shock front.
We have assumed a power law distribution of electrons with an isotropic
velocity distribution. Three alternative models have been considered for the
evolution of the magnetic field, including an equipartition model. We have
calculated the flux density at frequencies near the self absorption frequency
νa , under the assumption that νa ≪ νm .
We have obtained an expression for the self absorption frequency νa ,
which we defined as the frequency at the point where the extrapolations
of the asymptotic power laws above and below the νa meet. The value we
obtained for νa is close to the value obtained by Wijers & Galama (1998),
and a factor of ∼ 4 for p = 2.5 (or a factor of ∼ 7 for p = 2.2) larger than
the value obtained by Waxman (1997b).
We have calculated the observed spectra near νa for three different mag-
netic field models (Figure 4). The spectra differs more than a few percent
from the asymptotic power laws only for frequencies less than half an order
of magnitude above or below νa . This applies to all the magnetic field models
we have considered.
Together with the afterglow images obtained in GPS we now have a com-
plete set of the observed image over a wide range of frequencies (see Figures
4 and 5). For ν ≫ νm there is a thin bright ring on the outer edge of the
image, and the surface brightness at the center of the image is only a few
percent of its maximal value. For νa ≪ ν < νm there is a wider ring and the
contrast between the center and the edge of the image is smaller, with 58%
of the average surface brightness at the center. For ν ≪ νa the image is even
more homogeneous, with 77% of the average surface brightness at the center.
The observed image in the radio frequencies is of special importance, since
the best resolution is obtained in radio, with VLBI. A sufficient resolution
could be reached with a nearby GRB (z ∼ 0.2) to resolve the inner structure
of the image in radio frequencies.
We have fitted the theoretical spectra to observational data of the af-
terglow of GRB970508 (Shepherd 1998) and extracted the values of νa and
Fνa ,ext from the data (equation 22). Substituting these values into the the-
oretical expressions we obtained two equations for the physical parameters.
The behavior of the spectra near νm can supply two other equations, namely
the equations for the peak frequency and for the peak flux (adding only one
independent equation to the two equations for νa and Fνa ,ext). An additional
18
independent equation is obtained for νc . The power law p of the electron
distribution can be determined from the high energy slope (ν ≫ νm ).
It
is therefore possible with sufficient observational data to determine all the
physical parameters of the afterglow: p, ǫe , ǫB , E52 and n1 . A similar cal-
culation (using an equation for the cooling frequency νc instead of Fνa ,ext)
was made by Wijers & Galama (1998) for GRB970508. In fact combining
this equation we could even have an over-constrained system and check for
consistency of the solution. The equation we obtained from νa differs by a
factor of ∼ 2 from the corresponding equation given in Wijers & Galama.
This has a significant effect (up to a factor of ∼ 20) on the values of the phys-
ical parameters of the burst which they have given. We have shown (Table
1) that if other equations are corrected as well, the values of the physical
parameters vary even more (up to two orders of magnitude).
We thank Ehud Cohen for useful discussions and Dale Frail and Shri
Kulkarni for observational information. This research was supported by
NASA Grant NAG5-3516, and a US-Israel Grant 95-328. Re’em Sari thanks
The Clore Foundation for support.
References
[1] Blandford, R. D. & McKee, C. F. 1976, Phys. of fluids,19, 1130.
[2] Granot, J., Piran, T. & Sari, R. 1998. astro-ph/9806192.
[3] Gruzinov, A. & Waxman, E. 1998. astro-ph/9807111.
[4] Katz, J. 1994, ApJ, 422, 248.
[5] Katz, J. & Piran, T. 1997, ApJ, 490, 772.
[6] Kobayashi, S., Piran, T. & Sari, R. 1998. astro-ph/9803217.
[7] Panaitescu, A. & M´esz´aros, P. 1998, ApJ, 493, L31.
[8] Rybicki, G.B. & Lightman, A.P. 1979, Radiative Processes in Astro-
physics, Wiley-Interscience.
19
[9] Sari, R. 1997, ApJ, 489, L37
[10] Sari, R. 1998, ApJ, 494, L49
[11] Sari, R., Piran, T. & Narayan, R. 1998, ApJ, 497, L17.
[12] Shepherd, D.S., et al. 1998, ApJ, 497, 859.
[13] Taylor, D.A., et al. 1998, ApJ, 502, L115.
[14] Waxman, E. 1997a, ApJ, 485, L5.
[15] Waxman, E. 1997b, ApJ, 489, L33.
[16] Waxman, E. 1997c, ApJ, 491, L19.
[17] Wijers, R.A.M.J. & Galama, T.J. 1998, astro-ph/9805341.
[18] Wijers, R.A.M.J., Rees, M.J. &M´esz´aros, P. 1997, MNRAS, 228, L51.
20
|
astro-ph/0402337 | 1 | 0402 | 2004-02-13T16:28:36 | The stellar population associated with IRAS source 16132-5039 | [
"astro-ph"
] | We report the discovery of a young massive stellar cluster and infrared nebula in the direction of the CS molecular cloud associated to the IRAS point source 16132-5039. The analysis of the mid-infrared images from the more accurate MSX catalog, reveled that there are two independent components associated with the IRAS source. The integral of the spectral energy distribution for these components, between 8.28 $\mu$m and 100 $\mu$m, gave lower limits for the bolometric luminosity of the embedded objects of $8.7 \times 10^4 L_\odot$ and $9 \times 10^3 L_\odot$, which corresponds to ZAMS O8 and B0.5 stars, respectively. The number of Lyman continuum photons expected from the stars that lie along the reddening line for early-type stars is about 1.7 $ \times$ 10$^{49}$ s$^{-1}$, enough to produce the detected flux densities at 5 GHz. The NIR spectrum of the nebula increases with frequency, implying that free-free emission cannot be the main source of the extended luminosity, from which we conclude that the observed emission must be mainly dust scattered light. A comparison of the cluster described in this paper with the young stellar cluster associated with the IRAS source 16177-5018, which is located at the same distance and direction, shows that the mean visual absorption of the newly discovered cluster is about 10 magnitudes smaller and it contains less massive stars, suggesting that it was formed from a less massive molecular cloud. | astro-ph | astro-ph |
The stellar population associated with the IRAS source
16132-5039 1
A. Roman-Lopes, Z. Abraham
Instituto de Astronomia, Geof´ısica e Ciencias Atmosf´ericas, Universidade de Sao Paulo
Rua do Matao 1226, 05508-900, Sao Paulo, SP, Brazil
[email protected]
ABSTRACT
We report the discovery of a young massive stellar cluster and infrared nebula in the direction
of the CS molecular cloud associated to the IRAS point source 16132-5039. The analysis of the
mid-infrared images from the more accurate MSX catalog, reveled that there are two independent
components associated with the IRAS source. The integral of the spectral energy distribution for
these components, between 8.28 µm and 100 µm, gave lower limits for the bolometric luminosity
of the embedded objects of 8.7 × 104L⊙ and 9 × 103L⊙, which corresponds to ZAMS O8 and
B0.5 stars, respectively. The number of Lyman continuum photons expected from the stars that
lie along the reddening line for early-type stars is about 1.7 × 1049 s−1, enough to produce
the detected flux densities at 5 GHz. The NIR spectrum of the nebula increases with frequency,
implying that free-free emission cannot be the main source of the extended luminosity, from which
we conclude that the observed emission must be mainly dust scattered light. A comparison of
the cluster described in this paper with the young stellar cluster associated with the IRAS source
16177-5018, which is located at the same distance and direction, shows that the mean visual
absorption of the newly discovered cluster is about 10 magnitudes smaller and it contains less
massive stars, suggesting that it was formed from a less massive molecular cloud.
Subject headings: stars : formation -- stars: pre-main sequence -- infrared : stars -- ISM: HII regions --
ISM: dust, extintion
1.
Introduction
Massive stars are born within dense molecular
clouds forming clusters and associations; during
their formation and early evolution they are of-
ten visible only at infrared and radio wavelengths.
With the advent of the large bidimensional near-
infrared array detectors, the morphological and
photometric studies of extremely young galactic
stellar clusters have benefited. At near-infrared
wavelengths (1 to 2.5 µm) it is possible to probe
deep into the dense dust clouds where star forma-
tion is taking place.
The strong ultraviolet (UV) radiation emitted
by the massive stars dissociates and ionizes the
1Based on observations made at Laborat´orio Nacional
de Astrofisica/MCT, Brazil
gas, forming compact HII regions seen at radio
wavelengths. A large fraction of the radiation
could also heat the dust, which eventually radiates
in the far-infrared (FIR); for this reason, compact
HII regions are among the brightest and most lu-
minous objects in the Galaxy at 100µm.
Because massive stars evolve very fast, they are
extremely rare and difficult to find, except when
related to the emission of the surrounding molecu-
lar cloud. In that sense, CS and NH3 lines at radio
frequencies, characteristic of high density gas, are
good tracers of massive star forming regions.
In this paper we present observations in the
near infrared, of a young stellar cluster of mas-
sive stars in the direction of the IRAS point
source I16132-5039. This work is a part of a
survey aimed to the identification of stellar pop-
1
ulations in the direction of IRAS sources that
have colors characteristics of ultracompact HII
regions (Wood & Churchwell 1989) and strong
CS (2-1) line emission (Bronfman et al. 1996).
The studied region is located in the direction
of another massive star forming region asso-
ciated to the IRAS point source 16177-5018
(Roman-lopes et al. 2003); they are part of the
RCW 106 complex, located in the southern Galac-
tic plane at a distance of 3.7 kpc (Caswell & Haynes 1987).
The near-infrared cluster presented here had re-
cently been located by Dutra et al.
(2003) by
visual inspection of the 2MASS images; however
their work contains only the possible cluster loca-
tion, and as Persi, Tapia & Roth (2000) already
showed for NGC6334IV, a high concentration of
K band sources can be due to localized low ex-
tinction and be mistaken with a stellar cluster,
leading to false identifications.
The cluster is associated with the radio source
332.541-0.111, which presents continuum radio
emission at 5 GHz as well as hydrogen recombina-
tion lines (Caswell & Haynes 1987). The observa-
tions and data reduction are described in section
2, the results are presented in section 3 and our
main conclusions are sumarized in section 4.
2. Observations and data reduction
The imaging observations were performed in
June 2001 and May 2003 with the Near Infrared
Camera (CamIV) of Laborat´orio Nacional de As-
trofisica (LNA), Brazil, equipped with a Hawaii
1024x1024 pixel HgCdTe array detector mounted
on the 0.6 m Boller & Chivens telescope.The ob-
servations consisted of 8'x 8' frames in the direc-
tion of the IRAS source 16132-5039; the plate scale
was 0.47 arcsec/pixel and the mean values of the
PSF full width at half maximum (FWHM) were
1.4, 1.7 and 2.1 arcsec at the J , H and nbK im-
ages. The total integration time was 2220 s for
J , 1440 s for H and 6400 s for nbK filters, result-
ing in a sensitivity at 3σ of 18.2, 17.4 and 14.2
magnitudes and completeness limits of 17.8, 16.5
and 13.8 magnitudes respectively. Details about
the calibration and reduction procedures can be
found in Roman-Lopes et al. (2003).
Photometry from 2MASS All Sky Point Source
Catalogue 2 in the J , H and KS filters became
available recently, with completeness limits of
15.8, 15.1 and 14.3 magnitudes, in the three pass-
bands (Egan et al. 2001). Since the 2MASS KS
band photometry, centered at 2.17 µm and with
a bandpass of 0.32 µm has a completeness limit
greater than our nbK photometry, we completed
our observations with the 2MASS catalogue as-
trometry and magnitudes in this filter. The com-
parison of our photometry with the 2MASS survey
data, in an area of about 40 square arcmin, corre-
sponding to a total of 469, 610 and 440 common
sources in the J , H and K bands, respectively, is
shown in figure 1. We see a good linear relation
between the two systems, with a slope of 1 and a
dispersion that increases with the magnitude.
3. Results & Discussion
The combined false-colour infrared image, (J
blue, H green and nbK red) of the whole field is
displayed in Figure 2, together with amplified in-
dividual images of the nebular region at all bands.
All of them, but especially the H and nbK images,
shows the presence of a small spheroidal nebula
with a bright star at its center.
3.1. The IRAS source
In Figure 3 we present a contour map con-
structed from the LNA nbK image, with a beam
size of 2 × 2 pixels which shows the region around
IRAS16132-5039. The contours start at 2.2×10−4
Jy/beam and are spaced by this same value. The
IRAS coordinate has an intrinsic error delimited
by the ellipse plotted in the figure. A more accu-
rate position for the IR source was obtained from
the Midcourse Space Experiment - MSX point
source catalog (psc) 3. The MSX surveyed the
entire Galactic plane within b ≤ 5◦ in four mid-
infrared spectral bands centered at 8.28, 12.13,
14.65 and 21.34 µm, with image resolution of
19 arcsec and a global absolute astrometric ac-
curacy of about 1.9 arcsec (Price et al. 2001).
We found one MSX source, with coordinates
α(J2000) = 16h17m02.47s, δ(J2000) = −50d47m03.5s
within the IRAS uncertainty ellipse;
nates coincides with the star we labeled IRS1.
its coordi-
2http://www.ipac.caltech.edu/
3http://www.ipac.caltech.edu/ipac/msx/msx.html
2
However,
looking at the MSX images, we
found another closeby source, although outside
the IRAS uncertainty ellipse, with coordinates
α(J2000) = 16h16m55.94s, δ(J2000) = −50d47m07.8s.
In Figure 4 we present our H band image over-
laid with the 8.28µm band MSX contour diagram,
with the contours spaced by 8 × 10−6 W m−2
sr−1, starting at 2.8 × 10−5 W m−2 sr−1. The
same source was found in the contour plots from
the other MSX bands. We designated the stronger
source as A and the other as B. While the IRS1
object and the infrared nebulae are related to
source A , source B is also associated with a small
nebular region that shows a concentration of NIR
sources, as can also be seen in Figure 4.
Since only the 8.28 µm flux density was given
in the MSX psc for both sources, we integrated
the flux density of each individual source for the
four mid-infrared bands; the results are presented
in table 1, which also shows the values from MSX
and IRAS catalogs. Our integrated flux density
for source B in the 8.28 µm coincides within 5%
with the value given in the MSX catalogue, but
it is 30% larger for the A source. This discrep-
ancy can be understood if we consider that the
automatic MSX algorithm subtracted part of the
source B flux density as background contribution.
The relative contribution of source B decreases
with increasing wavelength, explaining also why it
was not resolved by the MSX algorithm. It should
be noticed that the reported IRAS flux density at
12 µm coincides with the sum of our derived flux
densities at 12.13 µm within 10%, while the MSX
value was about 50% lower.
In Figure 5 we plotted the mid to far-infrared
spectral energy distribution of the sources A and
B, without any correction for absorption. We as-
sumed that the IRAS flux density in the far in-
frared is divided between the two sources in the
same way as in the mid-infrared (14.65 and 21.33
µm): about 90% originating from source A and
10% from source B. We then integrated the ob-
served flux densities in the mid-far infrared, as-
suming a distance of 3.7 kpc, obtaining a lumi-
nosity LA = 8.7 × 104L⊙ and LB = 9 × 103L⊙
for sources A and B, respectively. Assuming that
the IR flux density represents a lower limit for
bolometric luminosity of the embedded stars, we
derived ZAMS spectral types O8 and B0.5 for
the sources A and B, respectively (Hanson et al.
3
1997).
3.2. Cluster population
In order to examine the nature of the stellar
population in the direction of the IRAS source, we
analyzed the stars in two delimited regions: one
that we labeled "cluster", which contains the neb-
ula and another that we labeled "control", which
has a stellar population that we believe is domi-
nated by "field" stars, as illustrated in Figure 6.
We represented the position of all objects detected
in the H band by crosses and we can see that the
small region labeled "cluster" shows a concentra-
tion of sources. In figure 7 we present comparative
(J − H) versus (H − K) diagrams for the stars de-
tected in the J, H and K images, together with
the position of the main sequence, giant branch
and reddening vectors for early and late type stars
(Koornneef 1983).
We see that the stellar population of the "clus-
ter" region is quite different from that of the "con-
trol" region, with many sources lying on the right
side of the reddening vector for early type stars,
showing excess emission at 2.2 µm. In the "con-
trol" region the majority of the sources have col-
ors of reddened photospheres, with many objects
located along the reddening vector for late type
stars. It is well established that very young pre-
main sequence objects present large infrared ex-
cess due to the presence of warm circumstellar
dust (Lada & Adams 1992). Our results suggest
that the stellar population in the direction of the
IRAS source is very young, as can be inferred from
their position in the (J − H) versus (H − K) dia-
gram.
We separated the cluster sources from the field
stars, by selecting all sources that lie to the right or
on the reddening vector for early type stars in the
cluster's color-color diagram. Table 2 shows the
coordinates and photometry of all selected sources
(J, H and K magnitudes from both LNA and
2MASS surveys).
Further information about the nature of the se-
lected objects in Table 2 can be extracted from
J versus (J − H) color-magnitude diagram shown
in Figure 8. We used this diagram instead the K
versus (H − K) color-magnitude diagram to min-
imize the efect of the "excess" of emission in the
NIR on the derived stellar spectral types. The lo-
cus of the main-sequence for class V stars at 3.7
kpc (Caswell & Haynes 1987) is also plotted, with
the position of each spectral type earlier than A0
V indicated. The intrinsic colors were taken from
Koornneef (1983) while the absolute J magnitudes
were calculated from the absolute visual luminos-
ity for ZAMS taken from Hanson et al. (1997).
The reddening vector for a ZAMS B0 V star, taken
from Rieke & Lebofsky (1985), is shown by the
dashed line with the positions of visual extinc-
tions AV = 10 and 20 magnitudes indicated by
filled circles. We also indicated the sources with
and without "excess" in the color-color diagram
by open and filled triangles respectivelly.
For the assumed distance, we estimated the
spectral type of the stars that do not present ex-
cess emission in the near infrared, by following the
de-reddening vector in the color-magnitude dia-
gram, for the others we only gave a rough classi-
fication; the results are shown in the last column
of Table 2. The main source of error in the de-
rived spectral types arises from uncertainties in
the assumed cluster distance, which was derived
from the velocities of the radio hydrogen recombi-
nation lines. Since the closest distance was used,
the main uncertainties come from the errors in the
galactic rotation curve model and related param-
eters. From the works of Blum et al. (1999, 2000,
2001) and Figueredo et al. (2002) comparing kine-
matic with spectroscopic distances we find that
they do not differ in more than 1 kpc, in which
case the change in the luminosity class would be
of two sub-spectral types for early O stars and one
for early B stars.
We must notice that source IRS1, which is as-
sociated with MSX source A, has an estimated
spectral type of at least O5, reddened by about
AV ≈ 14 magnitudes. Besides, there are five ob-
jects (IRS3, IRS11, IRS18, IRS21 and IRS33) as-
sociated with the mid-infrared source B; IRS3 is
probably an O8 ZAMS star reddened by about
AV = 7 magnitudes, while the others have esti-
mated spectral types of early-B stars. The corre-
sponding bolometric luminosities are in agreement
with the lower limits derived from the integrated
mid-far infrared flux densities, corresponding to
O8 and B0.5, for sources A and B respectively, as
seen in section 3.1.
A lower limit to the number of Lyman-continuum
photons produced in the star forming region, can
4
be calculated taking into account only the stars
that do not show "excess" in the color-magnitude
diagram (IRS1, 6, 8, 9, 10, 13, 14, 16, 20, 23, 26,
27 and 35 in Table 2). It was computed from the
relation given by Hanson et al. (1997), resulting
in 1.7 × 1049 photons s−1. Is interesting to note
that IRS1 is responsible for more than 90% of
the Lyman continuum photons, being the main
ionization source in the whole region.
It is also possible to obtain the number of ion-
izing Lyman continuum photons NLy from the ra-
dio continuum flux density given by Caswell &
Haynes (1987), using the expression derived by
Rubin (1968):
NLy =
5.59 × 1048S(ν)D2T −0.45
ν0.1
1 + fi[< He+/(H + + He+)]
e
(1)
where ν is in units of 5 GHz and fi is the fraction of
helium recombination photons that are energetic
enough to ionize hydrogen. Using the values of
Te = 4500 K, S(ν) = 3.3 Jy and D= 3.7 kpc taken
from Caswell & Haynes (1987), we find that fi ≈ 0
and NLy ∼ 6 × 1048 photons s−1, compatible with
the lower limit derived from the observed stars.
3.3. The Infrared Nebula
We can see from the detailed J, H and nbK
images in figure 2 that the nebula presents a
spheroidal shape, approximately symetric around
the IRS1 source.
In figure 9 we shown contour
maps of the nebular region at J , H and nbK bands
obtained from our infrared images. The contours
were calibrated in flux with the values starting
at 0.44 (J), 1.6 (H) and 2.2×10−4 Jy/beam (K),
with intervals of 0.37, 1.8 and 1.1×10−4 Jy/beam
respectively. We estimated the total flux density
by measuring the area between contours and mul-
tiplying by the value of the corresponding flux
density per unit area, obtaining S(J) = 0.023 Jy,
S(H) = 0.07 Jy and S(K) = 0.11 Jy. We then cor-
rected these results for extinction, using the mean
value < AV > = 15.1 magnitudes, derived from
the stars in the direction of the nebula that do not
present infrared excess (IRS1, IRS6 and IRS10),
and the standard extinction law from Rieke &
Lebofski (1985). We obtained S(J) = 1.16 Jy,
S(H) = 0.8 Jy and S(K) = 0.52 Jy.
In the previous section we showed that the num-
ber of ionizing photons available from the detected
stars is enough to explain the radio continuum flux
density measured by Caswell & Haynes (1987).
We will investigate now the contribution of free-
free emission to the observed nebular IR flux den-
sity. Assuming constant density and temperature
across the cloud and local thermodynamic equilib-
rium, the flux density S(ν) due to free-free emis-
sion can be written as:
S(ν) = τν Bν(T )Ω
(2)
where τν is the optical depth at frequency ν,
Bν(T ) is the Planck function
Bν(T ) =
2hν3
1
c2
exp(hν/kT ) − 1
(3)
and Ω is the solid angle of the source given by:
with even the less absorbed star (IRS1), we be-
lieve that this is not the main source of extended
IR emission.
Accepting the mean value of < AV > = 15.1
magnitudes as the mean value for the absorption
in the direction of A source, we found that the
corrected flux density increases with frequency,
suggesting that the observed extended radiation
is scattered light from the nearby stars. We ad-
justed then a black body to the NIR fluxes, ob-
taining a good fit for T≈ 16000K, characteristic
of middle-B stars (Hanson et al. 1997). Lumsden
& Puxley (1996), analyzing the ultracompact HII
region G45.12+0.13, also obtained an extinction
corrected flux density that increases with decreas-
ing wavelengths and interpreted it as due to stellar
light, scattered by dust through the HII region.
Ω = π(L/2D)2
(4)
3.4. Conclusions
where L is the diameter of the ionized cloud and
D the distance to the observer. The optical depth
at a given frequency ν is:
τν = αff L
(5)
where αff is the free-free absorption coefficient
(cgs) taken from Rybick (1979):
3.7 × 108 [1 − exp(−hν/kT )] neni gff (ν, T )
αff =
ν3 Z −2 T 1/2
(6)
where gff(ν, T ) is the Gaunt factor obtained from
Karzas & Latter (1961).
For two frequencies ν1 and ν2 the ratio of the
corresponding flux densities S(ν1) and S(ν2) may
be calculated from:
S(ν1)
S(ν2)
= exp[h(ν2 − ν1)/kT ]
gff(ν1, T )
gff(ν2, T )
(7)
For a flux density in the 5 GHz continuum of
3.3 Jy and an electron temperature Te = 4500 K,
as given by Caswell & Haynes (1987) the expected
flux densities at the J, H and K bands are 0.05,
0.10 and 0.16 Jy, respectively. We verify that the
measured values are much larger than what was
expected from the free-free emission, derived from
the radio data. In fact, only an absorption as low
as 4 magnitudes would explain the inferred spec-
trum as free-free emission. Since in the direction
of the A source, this absorption is incompatible
5
Near-IR imaging in the direction of the CS
molecular cloud associated with the IRAS source
16132-5039, revealed an embedded young massive
stellar cluster. We detected 35 member candidates
up to our completeness limit, concentrated in an
area of about 2 square parsec. All images, but
especially the H and nbK bands, show the pres-
ence of a small spheroidal nebula with a bright
star (IRS1) at its center.
The stars associated with the IRAS point
source were identified using more accurate po-
sitions from the MSX catalogue. The analysis of
the mid-infrared images reveled that there are two
sources associated with IRAS 16132-5039. The
strongest coincides with the position of at least
a dozen of OB stars, while the weaker source is
associated to less massive objects, with spectral
types characteristic of middle-B ZAMS stars. The
integral of the spectral energy distribution of the
MSX-IRAS sources, between 8.28 µm and 100
µm, gives lower limits to the bolometric luminos-
ity of the embeded objects of L = 8.7 × 104L⊙
and L = 9 × 103L⊙, which corresponds to ZAMS
O8 and B0.5 stars, respectively (Hanson et al.
1997). The results are compatible with the spec-
tral types of the objects detected in the NIR, since
it is possible that only part of the energy emitted
by the stars is reprocessed by the dusty envelope.
In that sense, they can be taken as lower limits to
the bolometric luminosity of the embeded stars.
Assuming that the radio emission measured by
Caswell & Haynes (1987) originates in this re-
gion, at a distance of 3.7 kpc, we estimated the
number of ionizing Lyman continuum photons as
NLy ∼ 6 × 1048 photons s−1. On the other
hand, the number of Lyman continuum photons
expected from the stars that lie along the redden-
ing line for early-type stars is about 1.7 × 1049
s−1, enough to produce the detected flux densities
at 5 GHz. The IRS1 source is enough to account
for more than 90% of the total number of Lyman
continuum photons necessary to ionize the gas.
Analysis of the integrated flux densities of the
NIR nebula at the J, H and nbK bands revealed
that they increase with frequency, implying that
free-free emission cannot be the main source of the
extended luminosity, unless we assume only four
magnitudes of visual extinction. Since this value
is incompatible with the extinction derived from
the stars that do not shown excess of emission at
2.2 µm, we conclude that the observed emission
must be mainly dust scattered light.
A comparison of the cluster described in this
paper with the young stellar cluster associated
with the IRAS source 16177-5018 (Roman-Lopes
et al. 2003), which is located at the same dis-
tance and direction, shows that the the former
contains less massive stars. Since its mean visual
absorption is also about 10 magnitudes smaller
than that of IRAS 16177-5018, it is possible that
it was formed from a less massive molecular cloud.
This work was partially supported by the
Brazilian agencies FAPESP and CNPq. We ac-
knowledge the staff of Laborat´orio Nacional de
Astrof´ısica for their efficient support and to An-
derson Caproni by help during the observations.
This publication makes use of data products
from the Two Micron All Sky Survey, which is
a joint project of the University of Massachusets
and the Infrared Processing and Analysis Cen-
ter/California Institute of Technology, funded by
the National Aeronautics and Space Administra-
tion and the National Science Foundation. This
research made use of data products from the Mid-
course Space Experiment.
REFERENCES
Blum, R. D., Damineli, A., Conti, P. S. 1999, AJ,
117, 1392
Blum, R. D., Conti, P. S., Damineli, A. 2000, AJ,
119, 1860
Blum, R. D., Damineli, A., Conti, P. S. 2001, AJ,
121, 3149
Bronfman, L. Nyman, L. A., May, J. 1996, A&AS,
115, 81
Caswell, J. L., Haynes, R. F. 1987, A&A, 171, 261
Dutra, C. M., Bica, E., Soares, J., Barbuy, B.
2003, A&A, 400, 533
Egan, M. P., Dyk, S. D. V., Price, S. D. 2001, AJ,
122,1844
Elias, J. H. 1982, AJ, 87, 1029
Figueredo, E., Blum, R. D., Damineli, A., Conti,
P. S. 2002, AJ, 124, 2739
Hanson, M. M., Howarth, I. D., Conti, P. S. 1997,
ApJ, 489, 718
Iben, I. Jr. 1965, ApJ, 141, 993I
Karzas, W. J., Latter, R. 1961, ApJ, 6, 167
Koornneef, J. 1983, A&A, 128, 84
Lada, C. J., Adams, F. C. 1992, ApJ, 393, 278
Lumsden, S. L., Puxley, P. J. 1996, MNRAS, 281,
493
Price, S. D., Egan, M. P., Carey, S. J., Mizuno, D.
R., Kuchar, T. A. 2001, AJ, 121, 2819
Rieke, G. H., Lebofsky, M. J. 1985, ApJ, 288, 618
Roman-Lopes, A., Abraham, Z., Caproni, A.,
L´epine, J. R. D. 2002, IAUS, 206, 244
Roman-Lopes, A., Abraham, Z., L´epine, J. R. D.
2003, AJ, 126, 1896
Rubin, R. H. 1968, ApJ, 154, 391R
Rybicki, G. B., Lightman, A. P. 1979, Radiative
Processes in Astrophysics, ed. JOHN WILEY
& SONS, New York
6
Simpson, J. P., Rubin, R. H. 1990, ApJ, 354, 165
Spitzer, L. 1978, Physical Processes in the Inter-
estellar Mediun, ed. Wiley, New York
Stetson, P. B., 1987, PASP, 99, 191
Persi, P. Tapia, M., Roth, M. 2000, A&A357, 1020
Wood, D. O. S., Churchwell, E. 1989, ApJ, 340,
265
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
7
Integrated fluxes from MSX A, C, D and E images for the two mid-infrared sources (see
text). Also are shown the data from MSX and IRAS point source catalogue (psc).
Table 1
λ(µm)
Integration "A"
Integration "B" MSX psc "A" MSX psc "B"
IRAS psc
8.28
12
12.13
14.65
21.33
25
60
100
15.8
· · ·
32.8
23.9
117
· · ·
· · ·
· · ·
4.3
· · ·
3.6
2.0
10.8
· · ·
· · ·
· · ·
11.3
· · ·
23.0
21.3
123.8
· · ·
· · ·
· · ·
4.1
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
44.8
· · ·
· · ·
· · ·
266
2311
4618
8
List of the selected near-infrared sources
Table 2
IRS
α(J2000)
δ(J2000)
JCamIV
J2mass
HCamIV
H2mass KCamIV
K2mass
Spec T ype
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
16:16:55.8
16:17:03.7
16:17:02.20 −50:47:03.1
16:17:09.23 −50:47:14.7
−50:47:23
16:17:00.62 −50:47:49.6
16:17:02.63 −50:46:56.6
−50:47:05
16:17:02.88 −50:47:05.1
16:17:06.30 −50:47:09.3
16:17:04.37 −50:47:16.4
16:17:03.72 −50:47:15.7
−50:47:22
16:16:59.72 −50:47:47.0
16:17:02.06 −50:47:26.6
16:17:08.61 −50:47:11.5
−50:47:01
16:17:04.16 −50:47:49.8
16:17:03.13 −50:47:15.4
−50:47:28
−50:47:08
16:17:02.02 −50:48:12.1
−50:47:25
−50:46:54
16:17:03.93 −50:47:34.0
−50:47:13
16:16:59.41 −50:47:52.9
16:17:08.12 −50:47:37.4
16:17:02.12 −50:47:37.2
−50:47:11
−50:47:35
16:17:03.88 −50:48:16.0
16:17:05.23 −50:47:31.7
16:17:06.74 −50:48:05.2
−50:47:20
−50:47:47
−50:47:03
16:16:55.9
16:17:03.8
16:17:03.5
16:16:55.6
16:17:03.7
16:16:55.3
16:17:02.5
16:17:03.9
16:16:56.0
16:17:02.0
16:17:03.8
16:17:02.2
12.26(4)
13.36(3)
11.77(2)
12.47(1)
14.43(3)
14.61(3)
17.08(9)
13.23(2)
13.28(2)
13.86(3)
13.12(3)
17.32(10)
16.27(6)
15.39(5)
16.23(6)
13.89(3)
14.74(3)
15.74(5)
15.53(5)
14.94(3)
16.22(8)
16.35(8)
16.50(7)
18.17(18)
17.16(9)
16.21(6)
16.54(8)
17.40(11)
16.63(7)
16.08(7)
15.28(4)
16.16(7)
13.62(5)
16.42(6)
15.67(5)
11.97(4)
13.26(4)
· · ·
· · ·
14.26(8)
· · ·
· · ·
13.31(3)
13.26(2)
13.92(4)
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
14.61(6)
· · ·
· · ·
14.90(5)
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
15.59(9)
15.20(9)
15.85(8)
· · ·
· · ·
· · ·
10.88(4)
11.66(2)
11.01(2)
12.02(2)
12.78(3)
13.04(2)
14.84(6)
12.28(3)
12.44(2)
12.57(3)
12.62(4)
14.76(6)
13.95(6)
13.76(3)
14.35(4)
13.27(3)
13.65(3)
14.24(4)
13.91(4)
13.89(4)
14.56(5)
14.47(5)
14.42(4)
15.16(7)
15.04(5)
14.47(4)
14.50(4)
15.36(6)
14.86(7)
15.54(7)
14.63(5)
15.37(6)
12.93(4)
14.25(5)
13.79(4)
10.65(4)
11.47(3)
· · ·
· · ·
12.51(7)
· · ·
· · ·
12.34(3)
12.45(3)
12.69(4)
· · ·
· · ·
· · ·
13.65(6)
· · ·
· · ·
13.57(5)
· · ·
· · ·
13.71(5)
· · ·
· · ·
14.42(8)
· · ·
· · ·
14.51(7)
· · ·
· · ·
· · ·
14.69(9)
· · ·
· · ·
· · ·
· · ·
· · ·
10.07(4)
10.13(3)
10.30(4)
11.17(4)
11.64(4)
12.03(5)
11.55(5)
11.42(3)
11.64(3)
11.80(4)
12.06
12.25(5)
12.54(5)
12.64(9)
12.51(6)
12.87(6)
12.71(6)
12.61(6)
12.69(7)
12.55(7)
12.95(6)
12.98(6)
13.31(7)
13.12(6)
13.11(6)
13.29(8)
13.42(7)
13.32(7)
13.61(7)
· · ·
13.53(5)
· · ·
11.90(4)
12.85(6)
12.56(6)
9.73(4)
10.15(3)
· · ·
10.92(2)
11.17(7)
· · ·
11.62(7)
11.70(3)
11.78(5)
11.92(5)
· · ·
12.31(6)
12.60(9)
12.65(2)
· · ·
12.72(7)
12.73(7)
· · ·
· · ·
12.88(5)
· · ·
· · ·
13.20(5)
· · ·
13.31(7)
13.38(6)
13.40(11)
· · ·
· · ·
13.74(8)
13.87(8)
14.18(10)
· · ·
· · ·
· · ·
O5
mid-O
early-B
early-B
early-B
B0
early-B
B0
B0.5
B0
early-B
early-B
O9.5
B0.5
early-B
B2
early-B
early-B
early-B
B2
early-B
early-B
B0.5
early-B
early-B
B2
B1
early-B
early-B
mid-B
· · ·
· · ·
early-B
B0
B0.5
∗Stars that show "excess" of emission at 2.2 µm
9
19
18
17
16
15
14
13
12
11
10
9
8
S
S
A
M
2
-
J
8
9
10
11
12
13
14
J-CamIV
15
16
17
18
19
16
15
14
13
12
11
10
9
8
S
S
A
M
2
-
H
16
15
14
13
12
11
10
9
8
7
6
S
S
A
M
2
-
K
6
7
8
9
10
11
12
13
14
15
16
K-CamIV
8
9
10
11
12
13
14
15
16
H-CamIV
Fig. 1. -- Magnitudes comparative diagram MJHK (2MASS) × MJHK (CamIV ). The continuous line shows
the relation expected if the two photometric systems were equal.
10
Fig. 2. -- Combined false-image made from the J, H and nbK LNA images. North is to the top east to the
right. At the botton are detailed views of the infrared nebulae J(lef t), H(center) and nbK(right) images.
11
40
50
-50:47:00
)
0
0
0
2
J
(
.
c
e
D
10
20
30
8
14
2
40
26
50
32
5
35
6
19
28
24
17
9
10
23
31
16
34
30
08
06
04
22
15
1
7
13
29
27
20
02
33
3
11
21
18
12
4
25
16:17:00
58
56
54
Right ascension (J2000)
Fig. 3. -- nbK band contour map of the infrared nebulae associated with IRAS 16132-5039. The contours
start at 2.2×10−4 Jy/beam, with the same intervals (the beam size is 2 × 2 pixels). The positions of selected
infrared sources refered in Table 2 and the IRAS coordinate elipse error (dotted line) also are indicated.
12
Fig. 4. -- Contour diagram from the A MSX band image (8.28µm), overlaying a LNA's H band image. The
contours start at 2.8×10−5 W/m2sr and are spaced by 8×10−6 W/m2sr.
13
-13.0
-13.5
-14.0
-14.5
-15.0
-15.5
-16.0
-16.5
-17.0
)
Fl
(l
g
o
L
MSX A
IRAS A
IRAS B
MSX B
0.9
1.0
1.1
1.2
1.3
1.4
1.5
Log (l )
1.6
1.7
1.8
1.9
2.0
2.1
Fig. 5. -- Spectral energy distribution of the "A" and "B" MSX sources. The mid infrared data (squares)
were obtained from the integrated flux of the individual sources in the MSX images (bands A=8.28µm,
C=12.13µm, D=14.65µm and E=21.34µm) while the far infrared data (triangles) were taken from IRAS,
and are listed in Table 1.
14
44m:24s
45m:36s
-50d:46m:48s
48m:00s
)
0
0
0
2
J
(
.
c
e
D
49m:12s
50m:24s
"cluster"
"control"
17m:24s
17m:12s
16h:17m:00s
16m:48s
R.A. (J2000)
Fig. 6. -- Diagram of the spacial distribution of sources detected at LNA's H band image. The two regions
(a) (cluster) and (b) (control) are delimitated by boxes .
15
4
3
2
1
0
H
-
J
AV= 10
10
AV= 0
"control"
30
30
20
20
4
3
2
1
0
H
-
J
30
"cluster"
30
24
12
25
15
28
21
18
7
AV= 20
AV= 10
27
29
14
6
13
34
22
23
35
2
5
19
10
8
9
16
3
31
11
26
1
20
17
30
33
4
32
AV= 0
0
1
2
3
0
1
2
3
H-K
H-K
Fig. 7. -- Color-color diagrams for two regions in our survey. The region labeled "control" (left panel)
contains only foreground objects; the other labeled "cluster" (right panel) has also a foreground population
(open circles) but presents objects with infrared "excess" (filled triangles). All objects with "excess" are
located in the "cluster area". The locus of the main sequence and giants branch are shown by the continuous
lines taken from Koornneef (1983), while the dashed and dotted lines follow the reddening vectors taken
from Rieke & Lebofsky (1985). The location (plus signs) of AV = 0, 10, 20 magnitudes of visual extinction
also are indicated. The cluster members candidates are labeled by numbers.
16
3
9
4
11
33
16
1
AV=10
8
10
O2v
O3v
O4v
O5v
O6v
O7v
O8v
O9.5v
B0v
B0.5v
B1v
B2v
B5v
A0v
J
8
10
12
14
16
18
2
5
17
20
6
AV=20
31
32
30
19
14
18
21
26
29
35
15
22
23
27
13
34
7
25
28
12
24
-0.5
0.0
0.5
1.0
1.5
2.0
2.5
3.0
J-H
Fig. 8. -- The J versus (J − H) color-magnitude diagram of the sources in table 2. The locus of the main
sequence at 3.7 kpc is shown by the continuous line. The intrinsic colors were taken from Koornneef (1983)
while the absolute J magnitudes were calculated from the absolute visual luminosity for ZAMS taken from
Hanson et al. (1997). The reddening vector for a B0 ZAMS star (dotted line) was taken from Rieke &
Lebofsky (1985). We also indicated the location (bold numbers) of AV = 10 and 20 magnitudes of visual
extinction as well as the sources that show "excess" (open up triangles) and do not (filled down triangles)
in the color-color diagram.
17
J
H
K
Fig. 9. -- The J, H and nbK contour maps of the infrare nebula region. The contours start at 0.44 (J), 1.6
(H) and 2.2×10−4 Jy/beam (K), and are spaced by 0.37, 1.8 and 1.1×10−4 Jy/beam respectively.
18
|
0704.1182 | 1 | 0704 | 2007-04-10T03:35:56 | An Optical Source Catalog of the North Ecliptic Pole Region | [
"astro-ph"
] | We present a five (u*,g',r',i',z') band optical photometry catalog of the sources in the North Ecliptic Pole (NEP) region based on deep observations made with MegaCam at CFHT. The source catalog covers about 2 square degree area centered at the NEP and reaches depths of about 26 mag for u*, g', r' bands, about 25 mag for i' band, and about 24 mag for z' band (4 sigma detection over an 1 arcsec aperture). The total number of cataloged sources brighter than r'= 23 mag is about 56,000 including both point sources and extended sources. From the investigation of photometric properties using the color-magnitude diagrams and color-color diagrams, we have found that the colors of extended sources are mostly (u*-r') < 3.0 and (g'-z') > 0.5. This can be used to separate the extended sources from the point sources reliably, even for the faint source domain where typical morphological classification schemes hardly work efficiently. We have derived an empirical color-redshift relation of the red sequence galaxies using the Sloan Digital Sky Survey data. By applying this relation to our photometry catalog and searching for any spatial overdensities, we have found two galaxy clusters and one nearby galaxy group. | astro-ph | astro-ph |
Accepted for Publication in ApJS
An Optical Source Catalog of the North Ecliptic Pole Region1
Narae Hwang2,11, Myung Gyoon Lee2, Hyung Mok Lee2, Myungshin Im2, Taehyun Kim2,
Hideo Matsuhara3, Takehiko Wada3, Shinki Oyabu3, Soojong Pak4, Moo-Young Chun5,
Hidenori Watarai6, Takao Nakagawa3, Chris Pearson3,7, Toshinobu Takagi3,
Hitoshi Hanami8, and Glenn J. White9,10
ABSTRACT
We present a five (u∗,g ′,r′,i′,z ′) band optical photometry catalog of the sources
in the North Ecliptic Pole (NEP) region based on deep observations made with
MegaCam at CFHT. The source catalog covers about 2 square degree area cen-
tered at the NEP and reaches depths of about 26 mag for u∗, g ′, r′ bands, about
25 mag for i′ band, and about 24 mag for z ′ band ( 4 σ detection over an 1
′′aperture). The total number of cataloged sources brighter than r′ = 23 mag
1Based on observations obtained with MegaPrime/MegaCam, a joint project of CFHT and
CEA/DAPNIA, at the Canada-France-Hawaii Telescope (CFHT) which is operated by the National Re-
search Council (NRC) of Canada, the Institut National des Science de l'Univers of the Centre National de
la Recherche Scientifique (CNRS) of France, and the University of Hawaii.
2Astronomy Program, Department of Physic and Astronomy, Seoul National University, Seoul 151-747,
Korea
3Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, Sagamihara, Kana-
gawa 229-8510, Japan
4Department of Astronomy and Space Science, Kyung Hee University, Yongin-si, Gyeonggi-do 446-701,
Korea
5Korea Astronomy and Space Science Institute, 61-1, Hwaam-Dong, Yuseong-Gu, Daejeon 305-348, Korea
6Office of Space Applications, Japan Aerospace Exploration Agency, Tsukuba, Ibaraki 305-8505, Japan
7ISO Data Centre, ESA, Villafranca del Castillo, Madrid, Spain
8Iwate University, 3-18-8 Ueda, Morioka 020-8550, Japan
9Astrophysics Group, Department of Physics, The Open University, Milton Keynes, MK7 6AA, UK
10Space Science & Technology Department, CCLRC Rutherford Appleton Laboratory, Chilton, Didcot,
Oxfordshire, OX11 0QX, UK
11e-mail: [email protected]
– 2 –
is about 56,000 including both point sources and extended sources. From the
investigation of photometric properties using the color-magnitude diagrams and
color-color diagrams, we have found that the colors of extended sources are mostly
(u∗ − r′) < 3.0 and (g ′ − z ′) > 0.5. This can be used to separate the extended
sources from the point sources reliably, even for the faint source domain where
typical morphological classification schemes hardly work efficiently. We have de-
rived an empirical color-redshift relation of the red sequence galaxies using the
Sloan Digital Sky Survey data. By applying this relation to our photometry
catalog and searching for any spatial overdensities, we have found two galaxy
clusters and one nearby galaxy group.
Subject headings: galaxies: general - galaxies: photometry - galaxies: clusters
- catalogs
1.
Introduction
The North Ecliptic Pole (NEP) is an undistinguished region in the sky, located at
α = 18h00m00s, δ = +66◦33′38′′. It is, however, a very special region since many astronomical
satellites have accumulated a large number of exposures over this location since Earth-
orbiting satellites must point their fixed solar panels to the Sun and be in their orbits over
the ecliptic poles. Unlike the South Ecliptic Pole (SEP) region where the South Atlantic
Anomaly and the LMC prevent the clear view of the extragalactic sky at certain wavelengths,
the NEP region suffers very little or no obscuration by foreground Galactic sources. The
galactic coordinates of the NEP are l ≈ 96.4◦, b ≈ +29.8◦ and the foreground extinction
in this direction is E(B − V ) = 0.047 (Schlegel et al. 1998). Therefore, the NEP is a
good target for deep, unbiased, contiguous surveys for extragalactic objects such as galaxies,
galaxy clusters, and AGNs.
The ROSAT All-Sky Survey (RASS) is one of the most representative surveys of the
NEP region (Voges et al. 1999). The ROSAT was the first X-ray imaging satellite to survey
the entire sky. A source catalog of X-ray sources that were extracted from a large area
surrounding the NEP was constructed from this RASS data (Henry et al. 2001, 2006). From
the followup observations and optical counterpart investigations, Gioia et al. (2003) reported
that most X-ray sources in the NEP region are AGNs (∼ 49%), stars (∼ 34%), and galaxy
groups/clusters (∼ 14%). Mullis et al. (2001) also found a supercluster of galaxies in the
NEP region using the RASS data and suggested that some galaxy clusters in this region are
part of the supercluster at z = 0.087. The NEP region was also observed in the 1.5 GHz
band by Kollgaard et al. (1994). Using this radio source catalog, Brinkmann et al. (1999)
– 3 –
investigated the correlation between the radio sources and the RASS X-ray sources, and
identified optical counterparts of the radio/X-ray sources. They found that a significant
number of radio loud sources are also bright in the X-ray band, and that X-ray selection is
an effective way for the search of galaxy clusters and groups.
While the X-ray luminosity is an efficient measure of the hot ionized gas in the galaxy
clusters, the star formation activities and the resultant stellar mass in galaxies can be esti-
mated from the optical and infrared flux. Recently, a new infrared space telescope, named
AKARI, was launched in February 2006. AKARI is expected to give exceedingly higher qual-
ity data than the previous infrared space missions such as Infrared Astronomical Satellite
(Neugebauer et al. 1984) and Infrared Space Observatory (Kessler et al. 1996). The AKARI
data will also complement the successful Spitzer Space Telescope (Werner et al. 2004) data
by providing better wavelength coverage over the 8 − 24µm band, which is not available from
the Spitzer. As one of its major science programs, AKARI is currently carrying out deep
near-to-mid infrared surveys over the wide area around the NEP through two major survey
programs, 'NEP-Deep' and 'NEP-Wide.' The key scientific objectives of these NEP surveys
are to unveil the dusty star formation history of the Universe, the mass assembly and large
scale structure evolution, and the nature of the cosmic infrared background (CIRB). For
further information on these NEP Programs, please refer to Matsuhara et al. (2006).
Combined with these space-borne infrared surveys, the optical survey of the NEP re-
gion constitutes the multifrequency dataset that is essential for the studies of cosmic star
formation history. Many researchers have investigated optical counterparts of radio and X-
ray sources in this region, but, for the imaging data, they mostly relied on the COSMOS
scan (e.g., Brinkmann et al. 1999) or Digitized Sky Survey images of the second Palomar
Observatory Sky Survey (POSS-II) plates, only complemented by some independent but
patchy observations (e.g., Gioia et al. 2003). Although the POSS-II data are excellent in
complete coverage of the whole sky, the usefulness of the data are still limited by the low
resolution with ∼ 1.0′′ per pixel scale and the shallow depth with the limiting magnitude of
rF ∼ 21.0 mag (Gal et al. 2004). In this study, therefore, we provide the first optical catalog
of sources in the NEP region based on deep observations in u∗, g ′, r′, i′, z ′ bands obtained
with MegaCam at CFHT. This optical catalog will be an important part of multifrequency
datasets of galactic and extragalactic sources in the NEP region. It will also serve as a basis
of the AKARI NEP survey mission by providing the information on the optical counterparts
of the infrared sources. This paper is composed as follows: We describe our observations in
Section 2 and data reduction procedures in Section 3. We present a bright source catalog
in Section 4.1, photometric properties of the sources in the NEP region in Section 4.2, and
galaxy number counts in Section 4.3. The result of galaxy cluster search using the color of
red sequence galaxies is given in Sections 4.4 and 4.5. Some photometric properties of other
– 4 –
X-ray sources are presented in Section 4.6 and our primary results are summarized in the
last section.
2. Observation
The observation of the North Ecliptic Pole (NEP) region was carried out over nine photo-
metric nights between August 22 and September 22, 2004 using the 3.6m CFHT telescope lo-
cated at Mauna Kea, Hawaii. We used a wide field imager MegaCam at the telescope primary
focus MegaPrime. The MegaCam is composed of 36 2048 × 4612 CCD's covering about 1◦ ×
1◦ area with 0.185 arc seconds resolution per pixel (Boulade et al. 2003). The u∗, g ′, r′, i′, z ′
filter system provided with MegaCam is basically the same as that used by the Sloan Digital
Sky Survey (SDSS) (York et al. 2000) except for the u∗ filter which is designed to take advan-
tage of the improved UV extinction condition at the CFHT site. Therefore, the photometric
data generated from this observation are presented in the CFHT unique u∗ and the SDSS
g ′, r′, i′, z ′ system. For further information on these filter systems used with MegaCam, see
the 'Filter set' section in http://www.cfht.hawaii.edu/Instruments/Imaging/MegaPrime/specsinformation.html
page.
As shown in Figure 1, the observed fields are composed of two fields separated by about
1◦ away from each other in the East-West direction, covering about two square degrees in
total. The NEP-E (East) field was observed with five filters, u∗, g ′, r′, i′, z ′ and the other
field, NEP-W (West), was observed with only four filters, g ′, r′, i′, z ′. Each field was covered
twice with a 15′′ offset for dithering in each filter with total exposure times of 3600 sec for
u∗ (1800 × 2), 2400 sec for g ′ (1200 × 2), 3000 sec for r′ (1500 × 2), 3000 sec for i′ (1500 ×
2), and 3600 sec for z ′ (1800 × 2). The raw images were processed using Elixir system by
CFHT staff. The Elixir is a collection of programs, databases, and other tools specialized
in processing and evaluation of the large mosaic data (Magnier & Cuillandre 2004). The
overall quality of the preprocessed images is good with a typical seeing of 0.7 ∼ 1.1′′. Detailed
information on the observation is listed in Table 1.
3. Data Reduction
3.1. Pre-Photometry Processing
For the reliable source detection and photometry, it is essential to prepare a master
image that satisfies the following two factors: (1) to obtain the required photometric depth
– 5 –
and (2) to give the complete areal coverage. We used a software package SWarp1 written
by E. Bertin at Terapix to transform each Multi Extension Fits (MEF) file containing 36
individual CCD frames into a single fits image data. SWarp also corrects different distortion
effects between input data through the resampling process, and then combines the resampled
images to create a deep output image after flux scaling and background subtraction.
To make a master image, we combined the g ′, r′, i′, z ′ four band data using SWarp for the
NEP-E and the NEP-W field respectively. During the SWarp run, BILINEAR resampling
method was used since this method was found to be effective in suppressing the discontinuity
effects around the chip boundaries, while other methods such as LANCZOS3 produced more
noticeable discontinuities. We also used weighted images during SWarp run to enhance the
quality of the output image. In the weight maps we assigned zero weights to pixels with
negative values. We excluded the u∗ band data from the master image construction since we
do not have the u∗ band data for the NEP-W and the S/N is relatively lower compared to
other bands. Finally, we created two master detection images with high S/N's, one for the
NEP-E and the other for the NEP-W region.
Two MEF images for each band were processed using SWarp with the same parame-
ters to generate a final photometry image for the corresponding filter. These u∗, g ′, r′, i′, z ′
band images were made to have the same dimension and coordinates with the master de-
tection image of the corresponding field. The depth of photometry attained by combin-
ing two raw images was calculated as 4σ flux over a circular aperture with 1 ′′ diameter.
The measured limiting magnitudes are u∗ ∼ 26.0, g ′ ∼ 26.1, r′ ∼ 25.6, i′ ∼ 24.7, and
z ′ ∼ 23.7 mag, as listed in Table 1. After all these processings, two master detection im-
ages and nine photometry images were prepared for the source detection and photometry.
All the information on the reduced photometry images will be available from an web site
http://astro.snu.ac.kr/ nhwang/index.files/nep.html.
3.2. Source Detection & Photometry
We have used SExtractor (Bertin & Arnouts 1996) to detect sources from the master
detection images of the NEP-E and NEP-W fields. A source is confirmed if it has more than
five contiguous pixels above four times the background sky fluctuation. The signals from each
source were measured in the u∗, g ′, r′, i′, z ′ band photometry images over the isophotal area
previously defined during the source detection. The photometry was made using SExtractor
in dual mode operation. This scheme enables the detection and photometry of any source
1See http://terapix.iap.fr/rubrique.php?id rubrique=49 for further information
– 6 –
that is registered at least once in any of the g ′, r′, i′, z ′ band images. The total number of
sources detected in our two observed NEP fields is about 130,000.
The instrumental magnitudes were transformed into standard magnitudes using the
transformation information provided and recorded in the header of the images by the CFHT
staff. During this transformation, only the calibrated magnitudes of sources that have avail-
able color information required for the calibration are calculated and kept in the catalog.
Otherwise, we assigned dummy values (99.000) to the magnitudes of sources in the final
source catalog.
4. Results
Figure 2 displays a source count histogram in the u∗, g ′, r′, i′, z ′ band along with the r′
band error and stellarity distributions. This shows that the number of sources in our data
increases up to u∗ ∼ 24.3, g ′ ∼ 24.0, r′ ∼ 23.5, i′ ∼ 23.0, and z ′ ∼ 22.2 mag before the
incompleteness effect starts to take place and the number of the detected sources starts to
decrease. The r′ band magnitude error is estimated to be about 0.1 mag or less at r′ ∼ 23.5
mag where the source count reaches its maximum. The stellarity distribution shows how
efficiently the star/galaxy classifier works in our dataset. From this distribution, it is clearly
seen that the stellarity index, which is calculated by SExtractor based on the isophotal areas,
peak intensity, and seeing information, separates the point sources (stellarity ∼ 1) and the
extended sources (stellarity ∼ 0) with high confidence for the sources with r′ < 22 mag. One
more point to be noted in this stellarity distribution is that sources with r′ < 16.5 mag and
stellarity > 0.7 are the results of the image saturation. Thus the stellarity index is most
relable in the magnitude range of 17 < r′ < 22 mag.
4.1. Source Catalog
Considering the source count and the stellarity distribution, we have decided the mag-
nitude range of the most reliable sources for the final bright source catalog entry, which is
r′ ≤ 23 mag. The number of sources compiled in the final bright source catalog is about
56,000. Table 2 lists a sample of the bright source catalog for reader's guide and the full
source catalog will be available electronically from the online Journal and from an web site
http://astro.snu.ac.kr/ nhwang/index.files/nep.html. Followings are the short descriptions
of data columns in the catalog.
Column 1 is the identification number of an optical source.
– 7 –
Column 2 & 3 list, respectively, the J2000.0 right ascension (RA) and declination (DEC)
of a source in degrees. The uncertainties of the astrometric solution derived by Elixir are
about ±0.5′′.
Column 4 ∼ 13 are the AB magnitudes and magnitude errors of a source in the
u∗, g ′, r′, i′, z ′ bands. These are given by MAG AUTO and MAGERR AUTO parameters
of SExtractor, which are Kron-like elliptical aperture magnitudes and their errors.
Column 14 is the extraction flag of a source given as a sum of powers of 2 by SExtractor.
If a source has neighbors (flag = 1) and is blended with another source (flag = 2) and some
pixels are saturated (flag = 4), then the final extraction flag given to the source is 7 (= 1
+ 4 + 2). Generally, a source with the extraction flag 0 gives the most reliable photometry.
For detailed information on this flag, see the SExtractor User's Manual.
Column 15 is the stellarity index of a source calculated from a g ′, r′, i′, z ′ combined
detection image. This index has a value between 1 (point sources) and 0 (extended sources).
See Figure 2 for the distribution of this index.
Column 16 is the ellipticity of a source calculated by SExtractor using the second order
image moments.
Column 17 is the FWHM in arcseconds of a source calculated under the assumption
that the source has a Gaussian profile.
Column 18 is the effective radius or a half-light radius of a source in arcseconds. This
value is computed by setting the input parameter PHOT FLUXFRAC = 0.5 and is given
by the output parameter FLUX RADIUS.
Column 19 is the semi-major axis of a source in arcseconds.
Column 20 is the detected isophotal area of a source. This parameter may be used as a
measure of the object's size for the case of extended sources.
Column 21 is the field number that the photometric data of a source comes from. The
field number 1 represents the NEP-E field and 2 represents the NEP-W field.
4.2. Color-Magnitude and Color-Color Diagrams
We have investigated the photometric properties of the NEP sources using several color-
magnitude and color-color diagrams. In each diagram, we use the stellarity index given by
SExtractor to distinguish between point sources and extended sources: stellarity > 0.95
for point sources (mostly stars) and stellarity < 0.2 for extended sources (mostly galaxies).
– 8 –
They are found to be statistically very useful tools to separate stars and galaxies brighter
than r = 23 mag from all kinds of sources in the catalog.
The r′ vs (g ′ − r′) color-magnitude diagram (CMD) and (g ′ − r′) color histogram in
Figure 3 shows two prominent vertical sequences of point sources, i.e., stars at (g ′ − r′) ∼
0.3 (G dwarf stars) and 1.4 (M giant stars), respectively. The extended sources, presumably
galaxies, are seen to be concentrated around the peak at (g ′ − r′) ∼ 0.8 that corresponds
to the (g ′ − r′) color of early type galaxies (Fukugita et al. 1995). The extended sources
start to dominate at r′ ≈ 22 mag and fainter. Some of these sources may be faint stars
that the SExtractor failed to classify as point sources. However, the different (g ′ − r′) color
histograms of the extended and the point sources suggest that the majority of these faint
extended sources are galaxies.
Figure 4 displays the r′ vs (u∗ − r′) CMD and (u∗ − r′) color histogram of sources in the
NEP-E field. The CMD shows that there are many galaxies distributed around (u∗ − r′) ∼
1.0 as shown in the (u∗ − r′) color histogram of extended sources. This value is consistent
with the (u∗ − r′) color of Scd type galaxies (Fukugita et al. 1995), whereas the (g ′ − r′) color
of most galaxies in Figure 3 is that of early type galaxies. It is also noted that there is a
very long redward tail in the (u∗ − r′) color distribution of extended sources, reaching nearly
(u∗ −r′) ≈ 4 or higher. The elliptical galaxies can be as red as (u∗ −r′) ∼ 2.8 (Fukugita et al.
1995). Therefore, some of those faint and red sources with (u∗ − r′) > 3.0 could be distant
galaxies with redshift z ≥ 0.1.
Figures 5 through 7 show the characteristic distribution of point sources and extended
sources in three color-color diagrams: (r′ − i′) vs (g ′ − r′), (i′ − z ′) vs (r′ − i′), and (g ′ − z ′)
vs (u∗ − r′). The most prominent feature in these color-color diagrams (CCD's) is very
distinguishable sequences of point sources, i.e. stars. In Figure 5, the stellar sequence in a
flipped 'L' shape has a very narrow width of about 0.2 dex in the (r′ − i′) vs (g ′ − r′) color
space. This tight sequence is also well represented by a straight line in Figure 6. In these
diagrams, extended sources are found to populate in a relatively well constrained color space
centered at a certain color. It is still clear that we can not possibly separate extended sources
from point sources by simply constraining colors in the (r′ − i′) vs (g ′ − r′) and the (i′ − z ′) vs
(r′ − i′) color-color space. However, Figure 7 shows that the (g ′ − z ′) vs (u∗ − r′) color-color
combination enables us to define the two exclusive spaces that are mostly populated by stars
and galaxies, respectively. The boundary of these regions can be drawn by combination
of three straight lines connecting (u∗ − r′, g ′ − z ′) = (−1.0, 0.5), (0.5, 0.5), (3.0, 1.7), and
(3.0, 5.0). This is a very useful tool for separating faint galaxies from stars to search for
distant galaxy clusters or to compute the 2D correlation function of galaxies.
– 9 –
4.3. Galaxy Number Counts
We have investigated the galaxy number counts using the NEP optical source catalog. To
select galaxies efficiently from the photometry catalog, we adopted two different definitions of
galaxies: (1) 'Galaxy I' sources with stellarity < 0.2 and (2) 'Galaxy II' sources that belong
to the designated area where the extended sources appeared to occupy in the (u∗ − r′) vs
(g ′ − z ′) CCD as denoted by the dashed line in Figure 7. The results of number counts for
Galaxy I and Galaxy II sample, as shown in Figure 8, show a consistent rise from r′ = 17.5
mag up to about r′ = 23 mag and then a steep down turn due to the incompleteness at
r′ ≥ 23 mag. From the lower and mid panels of Figure 8, it is clear that Galaxy II sample
outnumbers Galaxy I sample in r′ > 23 mag. This difference between number counts is
mostly due to the deteriorating reliability of stellarity index in r′ ≥ 23 mag, as shown in the
upper panel of Figure 2. Therefore, the Galaxy I sample is likely to lose a large number of
faint galaxies compared to the Galaxy II sample.
In Figure 8, we also plotted the R-band galaxy number count of Kummel & Wagner
(2001) for comparison after a simple transformation into r′ band. Kummel & Wagner (2001)
derived the number count of extended sources by modeling the number count distribution
of point-like sources (star) and then subtracting the extrapolated model count of point-like
sources from the detected source count. On the other hand, we defined galaxies as a certain
part of all the detected sources that satisfy a given stellarity condition (Galaxy I) or a
given two-color criterion (Galaxy II). Nonetheless, it is clear from the plot that the number
count by Kummel & Wagner (2001) is generally consistent with our result except for the
differences in the faint magnitude domain where the incompleteness becomes significant.
As shown in the mid panel of Figure 8, the logarithmic number counts defined as
Log(Count/0.5mag/deg2) also display very similar slopes over the range of r′ = 18 ∼ 22
mag. Simple linear square fits over that magnitude range returned the slope d(LogN)/dm =
0.387 for Galaxy I, 0.408 for Galaxy II, and 0.397 for Kummel & Wagner (2001) data. The
errors of the fitted slopes are about ±0.010 for all cases. Therefore, the slopes of logarithmic
galaxy number counts are in good agreement with that of Kummel & Wagner (2001) within
the errors.
However, there are some differences in several minor features between the logarithmic
profiles of galaxy number count. This is more clearly shown in the upper panel of Figure
8. It is apparent that there are two small dips in Galaxy II data: one at r′ ≈ 17.5 ∼ 18
mag and another smooth one at r′ ≈ 19 ∼ 20 mag. The first dip also appears to exist
in Kummel & Wagner (2001) data but the second smooth and shallow dip could not be
identified from Kummel & Wagner (2001) data in this plot. However, the magnitude of the
second dip is roughly coincident with the break point at R ≈ 19 ∼ 20 mag as shown in their
Figure 4 2.
– 10 –
4.4. Galaxy Clusters
We have carried out a galaxy cluster search using our optical photometry catalog.
Among numerous galaxy cluster finding methods available, we adopted a simplified version
of the 'C4 cluster finding algorithm' by Miller et al. (2005) that utilizes a seven-dimensional
position and color space. In this study, we use only a three-dimensional color space con-
structed based on (g ′ − r′), (r′ − i′), and (i′ − z ′) colors to select cluster member galaxies and
then we investigate the spatial distribution of the selected galaxies to find any overdensity
of these galaxies in small regions. Finally, the CMDs of any overdense region are consulted
to check whether the red sequence of galaxies is apparent before we identify the region as a
galaxy cluster.
This approach requires the definition of color ranges spanned by the cluster member
galaxies with various redshifts and richness classes before the actual application to the pho-
tometric data. We have used the Sloan Digital Sky Survey (SDSS) (York et al. 2000) data
for this purpose. The SDSS project provides a huge amount of multiband photometric and
spectroscopic data over a large sky area. This enables us to select cluster member galaxies
using the spectroscopy information and to reduce the possible contamination by field galax-
ies and stars. One more advantage of using the SDSS dataset to define color ranges occupied
by cluster member galaxies is that the SDSS filter system is the same as the MegaCam
filter system of CFHT except for the u∗ filter (see Section 2 for details). Therefore, we have
searched the SDSS database and retrieved photometric and spectroscopic data of nearby
galaxy clusters for the calibration of the red sequence colors in the SDSS filter system.
4.4.1. Nearby Galaxy Clusters in SDSS
We have searched the SDSS DR5 database (Adelman-McCarthy et al. 2007) and found
101 nearby Abell galaxy clusters whose u′, g ′, r′, i′, z ′ 3 photometric and spectroscopic data
2Figure 4 of Kummel & Wagner (2001) is constructed using the galaxy number count made with 0.25
mag bin. But the number count data in a table published in the same paper, which we used for Figure 8, is
made with 0.5 mag bin. The use of 0.5 mag bin instead of 0.25 mag bin in the number count is considered
to cause the smoothing out the feature.
3Please note that the u
′ filter is the Sloan system, which is different from the CFHT u
∗ system. See
Section 2 for details.
– 11 –
of member galaxies are available. Figure 9 displays the redshift distribution of these cluster
galaxies. The redshift of the selected galaxy clusters runs from 0 to 0.2 with a peak at
z ∼ 0.07. After the data retrieval from the SDSS data archive, we have selected member
galaxies of each cluster based on the velocity distribution and the spatial separation from the
cluster center. The total number of the selected member galaxies for the 101 clusters is about
5,700. From the cluster member galaxies, only the early type galaxies were selected using the
'f racDev' and 'eclass' parameters provided by the SDSS archive: f racDev > 0.8 and eclass
< 0 (for further information on these constraints, see Bernardi et al. 2005). Application
of these parameters finally returned about 1,700 early type galaxies in 88 clusters and the
number of early type galaxies in each cluster runs from 4 (A2192) to 73 (A2199) depending
on the redshift and the richness class.
Figure 10 displays the CMDs of the cluster galaxies in the SDSS color system. From
these diagrams, it is seen that most early type galaxies in the clusters lie within a well-
defined and narrow color range with a width of ≤ 0.2 dex in (g ′ − r′)0, (r′ − i′)0, and
(i′ − z ′)0. Although the (u′ − r′)0 vs r′
0 CMD shows a rather large dispersion, it still shows a
strong concentration at 2.0 ≤ (u′ − r′)0 ≤ 3.0. The colors of these narrow sequences of early
type galaxies have been used as indicators of the clusters' redshifts (Gladders & Yee 2000).
We used the SDSS photometry data of member galaxies in four clusters (A2199, A1166,
A1349, and A775) to derive the relation between the clusters' redshift and the (g ′ − r′)0 color
of the sequence as shown in Figure 11. From the linear fit after repeating one sigma clipping
twice, we derived a linear and empirical relation between the redshift and the (g ′ − r′)0 color
of the fitted sequences at r′
0 = 18 mag (hereafter (g ′ − r′)r18,0) as follows:
Redshift [z] = (0.415 ± 0.044) × (g ′ − r′)r18,0 − (0.286 ± 0.039)
(1)
Therefore, the galaxies in nearby clusters occupy a certain space in a multi-color parameter
space, which can be used to separate galaxies in the clusters from field galaxies and to
estimate the approximate redshift of the cluster using the (g ′ − r′)r18,0 color of the sequence.
4.4.2. Galaxy Cluster Search in the NEP Field
To find galaxy clusters based on the results shown in Figure 10 and to estimate the red-
shift using Equation 1, the Galactic foreground reddening was corrected for our photometry
data using Schlegel et al. (1998), assuming Ar′ ≈ 2.751E(B−V ), E(g ′−r′) ≈ 1.042E(B−V ),
E(r′ − i′) ≈ 0.665E(B − V ), and E(i′ − z ′) ≈ 0.607E(B − V ). After some tests on the CFHT
data, we have defined color ranges for the cluster galaxies as follows: 0.6 < (g ′ − r′)0 < 1.1,
0.1 < (r′ − i′)0 < 0.4, 0.0 < (i′ − z ′)0 < 0.4. We did not use (u∗ − r′)0 color for this color
– 12 –
space definition because of the unavailability of u∗ band data for the NEP-W field and the
general low S/N in the u∗ band in the NEP-E field. According to a test performed after
adopting the range 2.0 < (u∗ − r′)0 < 3.0 as the fourth constraining color, the efficiency of
finding the cluster galaxies turned out to be comparable to that of another test run using the
three color combination of (g ′ − r′)0, (r′ − i′)0, and (i′ − z ′)0. This three color combination
approach also enables a homogeneous search of cluster galaxies over our two data fields.
Possible cluster galaxies were selected by constraining the colors of galaxies with the
predefined parameters of (g ′ − r′)0, (r′ − i′)0, and (i′ − z ′)0. Using their RA and Dec in-
formation, we constructed spatial number density maps of the selected galaxies. Then we
identified 13 possibly overdense regions in the NEP field. Over these 13 regions, CMDs
of galaxies in (g ′ − r′)0, (r′ − i′)0, and (i′ − z ′)0 color were constructed using the source
catalog to find any feature that resembles the red sequence of galaxies in a cluster. From
this investigation, we have identified two galaxy clusters and one nearby galaxy group, and
estimated their redshifts using the (g ′ − r′)0 colors of the red sequences. The CMDs of those
galaxy clusters and the group are presented in Figures 12 − 14. More details about them
are discussed below.
4.5. Galaxy Clusters and Groups in X-ray/Radio Source Catalogs
Some galaxy clusters or groups were found and reported in the previous studies made
by using the X-ray and the radio band data of the NEP region. Henry et al. (1995) found
several X-ray-selected groups of galaxies in the NEP region based on RASS data and this
work was revised and extended further by Henry et al. (2006). Brinkmann et al. (1999) also
found that many X-ray sources in the NEP region have counterparts in the radio bands that
were observed with VLA. From the X-ray and radio source catalogs provided by Henry et al.
(2006) and Brinkmann et al. (1999), we have found a few galaxy clusters or groups in our
data field. Among these clusters and groups, we will discuss two galaxy clusters and a galaxy
group that were photometrically identified with our catalog.
4.5.1. NEPX1/VLA 1801.5+6645
This is the richest galaxy cluster found in our data field, which is easily seen in Figure 12.
In the upper panel of Figure 12, the CMDs of galaxies show very strong red sequences running
from r′
0 ≈ 16 mag (magnitude of the third brightest galaxy) in three color domains: (g ′ −r′)0,
(r′ − i′)0, and (i′ − z ′)0. Those galaxies belonging to the red sequences are concentrated in a
– 13 –
very compact region, as shown in the lower panel of Figure 12. The redshift estimated from
the red sequence's (g ′ − r′)r18,0 color and Equation 1 is about 0.072 ± 0.037. This cluster
was previously discovered by Burg et al. (1992) from ROSAT survey data and named as
NEPX1. They estimated the redshift of the cluster to be about 0.09 from the spectroscopic
observations, which is in good agreement with our estimate.
This cluster is considered as a part of the large supercluster structure that has been
reported to exist in the NEP region (Hasinger et al. 1991; Mullis et al. 2001). It was also
detected in the radio band from the VLA observations and was suggested as a possible coun-
terpart of the X-ray source RXS J180137.7+664526 by Brinkmann et al. (1999). Although
Brinkmann et al. (1999) listed this source as a galaxy group, we have reached a conclusion
based on the red sequence galaxies in the CMDs that it is a galaxy cluster rather than a
galaxy group.
4.5.2. RX J1754.7+6623
Henry et al. (2006) classified this source as a galaxy cluster with redshift z = 0.0879.
However, the CMDs shown in Figure 13 suggest that there are two kinds of galaxies: (1)
several bright galaxies that form the red sequence running from r′
0 ≈ 16 to 18 mag and (2)
many faint galaxies in the background. The second component of galaxies are not brighter
0 = 19 mag, which is about 1 mag fainter than the faintest galaxy of the brighter
than r′
component. This indicates that those bright galaxies may happen to be located in front
of the faint background galaxies by chance. Therefore, we suggest that these galaxies are
members of a galaxy group rather than a galaxy cluster. The red sequence's (g ′ − r′)r18,0
color is about 0.793, corresponding to an estimated redshift of 0.043 ± 0.048. The relative
large error may be due to the poorly determined slope of red sequence depends sensitively
on several bright galaxies with r′
0 < 18 mag.
4.5.3. RX J1757.9+6609
There is a known X-ray source RX J1757.9+6609 at the similar position. It is listed as
a type 2 AGN with redshift z = 0.4865 in the catalog of Henry et al. (2006). A type 2 AGN
is a narrow emission line (FWHM < 2000 km s−1) object, while a type 1 AGN is a broad
emission line (FWHM ≥ 2000 km s−1) object (Gioia et al. 2003). The spatial number density
plot (lower panel) of the red sequence galaxies in Figure 14 shows some clustering in the 2D
projected space. This point was also noted by Gal et al. (2003) who carried out a cluster
– 14 –
search using the digitized Second Palomar Observatory Sky Survey (DPOSS) data. They
detected an overdensity of galaxies at this location and listed it as a galaxy cluster candidate
under the name of NSC J175751+660924, and also estimated its redshift to be z ≈ 0.1663
from their photometric redshift analysis. Our photometric data shows well developed red
sequences running from r′
0 ≈ 17.5 mag in (g ′ − r′)0, (r′ − i′)0, and (i′ − z ′)0 colors, confirming
that it is a genuine galaxy cluster. The red sequence's (g ′ − r′)r18,0 color is about 0.811 and
this leads to the estimated redshift of z = 0.043 ± 0.025. Therefore, there appears to be
a galaxy cluster with redshift z < 0.2 and a type 2 AGN with redshift z = 0.4875 in the
background of the same projected area.
4.5.4. Other Galaxy Clusters
We have identified two galaxy clusters and one galaxy group by applying the simple
color cut method to our photometry catalog. However, there are two more galaxy clusters
that are listed in the literature in our observation field that we failed to confirm. One is RX
J1801.7+6637 (z = 0.57) reported by Bower et al. (1996) and the other is RX J1757.3+6631
(z = 0.6909) from the catalog of Henry et al. (2006). There appear to be some weak hints
of clustering around these clusters on the images. But our photometry is not deep enough
to identify any red sequence of galaxies in the CMD of the galaxies.
4.6. The Photometric Properties of Other X-ray Sources
The photometric properties of various X-ray sources such as stars, AGNs, BL Lac's could
provide valuable information regarding their stellar populations and evolution. Comparison
of our photometry catalog with the X-ray source catalog of Henry et al. (2006) returned
about 25 possible matches. Based on the classification information of the X-ray sources by
Henry et al. (2006), we have investigated the photometric property of each object class using
several CMDs and CCDs. Figure 15 shows a (g ′ − r′) vs r′ CMD and the three different
CCDs of various optical counterparts of X-ray sources. For a guideline in each diagram, the
distribution of point sources, which are mostly stars, is also plotted.
In Figure 15, it is easily seen that (1) generally, X-ray bright stars do not belong to the
well defined stellar sequence in each diagram, and (2) AGNs and BL Lac objects are not
readily separated from stars or other kind of objects in these CMDs and CCDs. Most X-ray
bright stars that are used in this comparison are very bright (r′ < 15 mag) and are mostly
saturated except for one faint giant star as shown in the upper-left panel of Figure 15. This
– 15 –
may explain why these X-ray bright stars appear to be in different color domains from other
generic stars. Even one faint star (marked by a star with a circle in Figure 15) is bluer in
(u∗ − r′) color than normal stars. Although AGNs and BL Lac's as a whole do not show any
distinct pattern in each color, the AGN1s with stellarity > 0.9, which are relatively free from
wrong identification, are found only in the blue domains of (g ′ − r′), (r′ − i′), and (u∗ − r′).
The (g ′ − r′) CMD also shows that any sources with (g ′ − r′) < 0.0 and r′ > 15 ∼ 16 mag
are very likely to be AGN or BL Lac objects.
5. Summary and Conclusion
We have obtained u∗, g ′, r′, i′, z ′ optical band high resolution images of the two square
degrees area centered at the NEP with MegaCam/MegaPrime at CFHT. From the source
detection and the photometry using SExtractor, we have compiled about 56,000 sources with
r′ ≤ 23 mag, including point and extended sources, into the final optical photometry catalog
as listed in Table 2. The use of the color-magnitude diagrams and color-color diagrams
revealed strikingly different photometric characteristics of stars and galaxies. We have found
that the use of (u∗ − r′) vs (g ′ − z ′) color enables us to clearly separate galaxies from stars
and this separation does not suffer from the uncertainties involved in the morphological
classification of faint sources. The galaxy number counts constructed from the galaxies
selected based on the (u∗ − r′) vs (g ′ − z ′) color show a nearly monotonic increase up to
about r′ = 23 mag with a slope d(LogN)/dm ≈ 0.40 ± 0.01, which is in agreement with
the literature. However, there are some changes in the slope at r′ = 17.5 ∼ 18 mag and
r′ = 19 ∼ 20 mag, which needs further studies.
Using the SDSS DR5 data of the 101 nearby Abell galaxy clusters with redshift z < 0.2,
we have derived a relation between the redshift and the color of the red sequences in the
SDSS filter system which is compatible with the CFHT MegaCam filter system. Utilizing
the information derived from the nearby galaxy clusters, we have applied a simple color
cut method to find galaxy clusters in our data field, which returned two galaxy clusters
and one galaxy group. These galaxy clusters and group are also radio and X-ray sources,
which were reported by previous studies. We have also estimated the redshift of these galaxy
clusters and group using the linear relation between the (g ′ − r′)r18,0 color and the redshift.
The estimated redshift is in agreement with the known value of 0.09 for galaxies in NEPX1
but it is relatively lower than the spectroscopic redshift in the literature for galaxies in
RX J1754.7+6623 and the photometrically derived redshift for RX J1757.9+6609. For RX
J1754.7+6623, this underestimation of redshift may be due to the poorly determined slope
of red sequence since there are only a few bright galaxies. For RX J1757.9+6609, the galaxy
– 16 –
cluster seems to be overlayed on the background type 2 AGN with redshift of 0.4875.
We have compared our photometry catalog with an X-ray source catalog in the literature
to investigate the photometric properties of other X-ray sources. The result of comparison
implies that the sources with (g ′ − r′) < 0.0 could be classified as candidates of AGNs or BL
Lac objects.
N.H. and T.K. are in part supported by the BK21 program of the Korean Government.
This work was in part supported by the ABRL (R14-2002-058-01000-0).
Adelman-McCarthy, J., et al. 2007, submitted to ApJS
REFERENCES
Bernardi, M., Sheth, R. K., Nichol, R. C., Schneider, D. P., & Brinkmann, J. 2005, AJ, 129,
61
Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
Brinkmann, W., et al. 1999, A&AS, 134, 221
Burg, R., et al. 1992, A&A, 259, L9
Boulade, O., et al. 2003, SPIE, 4841, 72
Bower, R. G., et al. 1996, MNRAS, 281, 59
Fukugita, M., Shimasaku, K., & Ichikawa, T. 1995, PASP, 107, 945
Gal, R. R., et al. 2003, ApJ, 125, 2064
Gal, R. R., et al. 2004, AJ, 128, 3082
Gioia, I. M., et al. 2003, ApJS, 149, 29
Gladders, M. D., & Yee, H. K. C. 2000, AJ, 120, 2148
Hasinger, G., Schmidt, M., & Trumper, J. 1991, A&A, 246, L2
Henry, J. P., et al. 1995, ApJ, 449, 422
Henry, J. P., et al. 2001, ApJ, 553, L113
– 17 –
Henry, J. P., et al. 2006, ApJS, 162, 304
Kessler, M. F., et al. 1996, A&A, 315, L27
Kollgaard, R. I., et al. 1994, ApJS, 93, 145
Kummel, M. W., & Wagner, S. J. 2001, A&A, 370, 384
Magnier, E. A., & Cuillandre, J.-C. 2004, PASP, 116, 449
Matsuhara, H., et al. 2006, PASJ, 58, 673
Miller, C. J., et al. 2005, AJ, 130, 968
Mullis, C. R., et al. 2001, ApJ, 553, L115
Neugebauer, G., et al. 1984, ApJ, 278, L1
Schlegel, D., Finkbeiner, D., & Davis, M., 1998, ApJ, 500, 525
Voges, W., et al. 1999, A&A, 349, 389
Werner, M. W., et al. 2004, ApJS, 154, 1
York, D. G., et al. 2000, AJ, 120, 1579
This preprint was prepared with the AAS LATEX macros v5.2.
– 18 –
Table 1. Observation Log
Field
RA
DEC
Filter
Total Exp Time
(hh:mm:ss)
(dd:mm:ss)
NEP-E
18:04:31.51
66:33:38.60
NEP-W 17:55:28.49
66:33:38.60
(sec)
1800 × 2
1200 × 2
1500 × 2
1500 × 2
1800 × 2
1200 × 2
1500 × 2
1500 × 2
1800 × 2
u∗
g′
r′
i′
z ′
g′
r′
i′
z ′
Seeing
(arcsec)
Depth
Observation Date (UT)
(AB mag)a
(yyyy-mm-dd)
1.13
1.08
0.99
0.69
0.73
1.01
0.87
0.99
0.85
25.98
26.12
25.58
24.70
24.03
26.12
25.58
24.85
23.71
2004-09-13/14
2004-08-22
2004-08-23
2004-08-23/09-12
2004-09-07
2004-08-22
2004-09-19/20
2004-09-13/14
2004-09-22
aThe depth of each filter data was measured as 4 σ flux over a circular aperture with a diameter of 1 ′′.
ID RA (J2000) Dec (J2000)
u∗
err(u∗)
g′
err(g′)
r′
err(r′)
i′
err(i′)
z′
err(z′) flag stellarity ellipticity
Table 2. NEP Optical Source Cataloga
[deg]
[deg]
[mag]
[mag]
[mag]
3 271.1438293
5 271.4050293
6 272.0104980
17 270.6894836
20 270.6847839
24 271.0955811
25 270.9777527
26 270.6968384
35 271.8403320
37 270.7434998
66.0613251
66.0613251
66.0599136
66.0615005
66.0614319
66.0624542
66.0611649
66.0618744
66.0610962
66.0618744
0.098
0.033
99.000 99.000 25.438
22.822
99.000 99.000 24.105
22.895
99.000 99.000 99.000 99.000 18.853
22.994
99.000 99.000 24.819
22.875
99.000 99.000 26.160
22.803
23.651
23.783
19.145
99.000 99.000 20.694
23.781
25.055
22.911
22.855
24.433
24.490
99.000 99.000 24.161
22.042
0.066
0.182
0.026
0.002
0.061
0.041
0.036
0.040
0.050
0.072
[mag]
22.608
22.200
18.484
21.750
21.472
22.544
18.382
22.476
22.677
20.852
0.031
0.019
0.000
0.014
0.010
0.028
0.001
0.022
0.028
0.006
[mag]
26.193
1.738
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
99.000 99.000
0.020
0.020
0.001
0.025
0.022
0.019
0.001
0.018
0.018
0.011
0
0
4
0
0
0
0
0
0
0
fwhm
[arcsec]
r eff
sma
area
field
[arcsec]
[arcsec]
[arcsec2]
0.08
0.76
1.00
0.28
0.18
0.76
0.99
0.88
0.88
0.98
0.295
0.042
0.162
0.056
0.133
0.110
0.132
0.129
0.160
0.090
1.611
0.990
0.773
1.373
2.218
1.117
1.025
1.175
1.032
0.995
0.620
0.460
0.326
0.562
0.580
0.531
0.407
0.484
0.468
0.506
0.423
0.347
0.156
0.389
0.398
0.334
0.481
0.318
0.333
0.408
1.677
1.814
0.274
1.814
1.882
1.437
6.640
1.369
1.506
2.772
1
1
1
1
1
1
1
1
1
1
aThe complete version of this table is in the electronic edition of the Journal. The printed edition contains only a sample.
–
1
9
–
– 20 –
Fig. 1.- The observation field map around the North Ecliptic Pole (NEP) (dashed line
boxes). Squares and asterisks represent, respectively, the known radio (Brinkmann et al.
1999) and X-ray (Henry et al. 2006) sources. Circles with an asterisk represent galaxy groups
or clusters classified by Gioia et al. (2003) and cataloged in Henry et al. (2006). The large
cross at the center shows the location of NEP. There are about 40 radio and/or X-ray sources
found in our two fields of observation, NEP-E (East) and NEP-W (West).
– 21 –
Fig. 2.- The distribution of the stellarity index (upper panel), the r′ band magnitude error
(mid panel), and the source number counts in u∗, g ′, r′, i′, z ′ bands (lower panel). The number
of detected sources reaches its maximum at about 24.3 mag for u∗, 24.0 mag for g ′, 23.7 mag
for r′, 23.0 mag for i′, and 22.5 mag for z ′ band.
– 22 –
Fig. 3.- Upper panel: (g ′ − r′) vs r′ color magnitude diagram (CMD) of NEP sources.
Sources selected with r′ ≤ 23 mag and (g ′ − r′) color error < 0.1 mag criteria are plotted
in different colors: point sources with stellarity > 0.95 in red and extended sources with
stellarity < 0.2 in blue. Black dots with r′ < 17 mag mostly represent saturated stars.
Lower panel: (g ′ − r′) color distribution of those selected sources: point sources in solid
line and extended sources in dashed line. Point sources and extended sources show different
distributions in (g ′ − r′) color: two peaks at (g ′ − r′) ≈ 0.3 (G dwarf stars) and 1.4 (M giant
stars) for point sources and a single peak at (g ′ − r′) ≈ 0.8 for extended sources.
– 23 –
Fig. 4.- Upper panel: (u∗ −r′) vs r′ color magnitude diagram (CMD) of sources in the NEP-
E field. Sources selected with r′ ≤ 23 mag and (u∗ − r′) color error < 0.1 mag criteria are
plotted in different colors: point sources with stellarity > 0.95 in red and extended sources
with stellarity < 0.2 in blue. Lower panel: (u∗ − r′) color distribution of those selected
sources: point sources in solid line and extended sources in dashed line.
– 24 –
Fig. 5.- Panel A: (g ′ − r′) vs (r′ − i′) color-color diagram (CCD) of NEP sources. Sources
selected with r′ ≤ 23 mag, (g ′ − r′) color error < 0.1 mag, and (r′ − i′) color error < 0.1
mag criteria are plotted: point sources with stellarity > 0.95 in crosses and extended sources
with stellarity < 0.2 in dots. The color distribution in (g ′ − r′) color (Panel B) and (r′ − i′)
color (Panel C) of those selected sources: point sources in solid line and extended sources in
dashed line. Please note that the (r′ − i′) and (g ′ − r′) color sequence of point sources is very
narrow and clear while the feature of extended sources is very diffuse and broad centered at
about (g ′ − r′) ∼ 0.8 and (r′ − i′) ∼ 0.3.
– 25 –
Fig. 6.- Panel A: (r′ − i′) vs (i′ − z ′) color-color diagram (CCD) of NEP sources. Sources
selected with r′ ≤ 23 mag, (r′ − i′) color error < 0.1 mag, and (i′ − z ′) color error < 0.1
mag criteria are plotted: point sources with stellarity > 0.95 in crosses and extended sources
with stellarity < 0.2 in dots. The color distribution in (r′ − i′) color (Panel B) and (i′ − z ′)
color (Panel C) of those selected sources: point sources in solid line and extended sources
in dashed line. In this plot, point sources are distributed in a narrow straight line running
from (r′ − i′, i′ − z ′) ≈ (0.1, 0.1) to (1.8, 0.8).
– 26 –
Fig. 7.- Panel A: (u∗ − r′) vs (g ′ − z ′) color-color diagram (CCD) of NEP sources in the
NEP-E field. Selected sources with r′ ≤ 23 mag, (u∗ − r′) color error < 0.1 mag, and (g ′ − z ′)
color error < 0.1 mag criteria are plotted: point sources with stellarity > 0.95 in crosses and
extended sources with stellarity < 0.2 in dots. The dashed line in Panel A separates point
sources from extended sources. See text for details. The color distribution in (u∗ − r′) color
(Panel B) and (g ′ − z ′) color (Panel C) of those selected sources: point sources in solid line
and extended sources in dashed line.
– 27 –
Fig. 8.- The number counts of galaxies in the NEP region, normalized to one square
degree area. The Kummel & Wagner (2001) (KW01) data in R band are also shown in all
the three panels for comparison with our data (Galaxy I & II) assuming a simple relation
of r′ ≃ R + 0.24. Galaxy I is defined to be sources with stellarity < 0.2 while Galaxy
II is selected based on its position on the (u∗ − r′) vs (g ′ − z ′) CCD. It is clear that the
overall shapes and patterns of galaxy number counts are generally in good agreement with
Kummel & Wagner (2001) result.
– 28 –
Fig. 9.- The redshift distribution of 101 galaxy clusters whose photometric and spectro-
scopic data of member galaxies were retrieved from the SDSS DR5 data archive.
– 29 –
Fig. 10.- The color-magnitude diagrams (CMD) of galaxies in 101 nearby (z < 0.2) galaxy
clusters in the SDSS data archive. All the photometric data used in these diagrams are
foreground reddening corrected. Early type galaxies are plotted in crosses. Please note the
relatively narrow and well-defined red sequence of early type galaxies in each CMD.
– 30 –
Fig. 11.- The representative (g ′ − r′)0 color red sequences of four galaxy clusters: A2199
(inverted triangle, z = 0.0302), A1169 (square, z = 0.0586), A1346 (triangle, z = 0.0975),
and A775 (circle, z = 0.1334). The color of the red sequence was derived by fitting the
colors of early type galaxies in each cluster marked by the filled symbols while the open
symbols were rejected from the fit. The red sequence of a galaxy cluster shifts redward as
the corresponding redshift increases.
Fig. 12.- The CMDs (the three upper panels), the spatial distribution (lower left), and the
number density plot (lower right) of galaxies of NEPX1. In each CMD, only the extended
sources are plotted and the magnitude limit adopted for the analysis is indicated with an
arrow. The color ranges used to find the cluster galaxies are shown in short dashed lines and
the red sequence line fitted in each color is displayed in long dashed lines. The estimated
redshift from the (g ′ − r′) color of the red sequence is indicated in the head as 'ZCMR'. The
galaxies satisfying the color cut criteria and being located within a solid circle in the lower
left panel are represented by filled circles in the CMDs and the spatial distribution plot.
The size of the circles in the spatial distribution plot are proportional to luminosity: the
larger, the brighter. Galaxies that satisfies only the color cut criteria are plotted in open
circles while other galaxies are in open squares. The dashed line in the density plot displays
the number density profile of all galaxies around the selected area and the solid line shows
the profile of ten times the number density of those selected galaxies and the corresponding
errors.
– 32 –
Fig. 13.- The CMDs (the three upper panels), the spatial distribution (lower left), and the
number density plot (lower right) of galaxies of RX J1754.7+6623. There are several bright
galaxies in the foreground over the many faint background galaxies in this region. See Figure
12 for the legend.
– 33 –
Fig. 14.- The CMDs (the three upper panels), the spatial distribution (lower left), and the
number density plot (lower right) of galaxies of RX J1757.9+6609. See Figure 12 for the
legend.
– 34 –
Fig. 15.- The photometric properties of X-ray sources. Stellar symbols for stars, triangles
for type 1 AGN (AGN1), reversed triangles for type 2 AGN (AGN2), asterisks for BL Lac's,
squares for galaxies, filled circles for galaxy clusters, and dots for point sources from our
photometry catalog. The classification information of X-ray sources are from Henry et al.
(2006). The distribution of point sources is presented in each CMD or CCD for references.
Stars marked by stellar symbols in these diagrams are suspected to suffer from saturation
effect except for one faint star represented by a stellar symbol within a circle. Please note
that the point-like AGN1 sources (filled triangles) are bluer than (g ′ − r′) ∼ 0.5.
|
astro-ph/0003119 | 1 | 0003 | 2000-03-08T22:41:34 | VSOP Observations of a Sub-Parsec Accretion Disk | [
"astro-ph"
] | The physical conditions in the inner parsec of accretion disks believed to orbit the central black holes in active galactic nuclei can be probed by imaging the absorption of background radio emission by ionized gas in the disk. We report high angular resolution observations of the nearby galaxy NGC 4261 which show evidence for free-free absorption by a thin, nearly edge-on disk at several frequencies. The angular width, and probably the depth, of the absorption appears to increase with decreasing frequency, as expected. Because free-free absorption is much larger at lower frequencies, the longest possible baselines are needed to provide adequate angular resolution; observing at higher frequencies to improve resolution will not help. | astro-ph | astro-ph |
100
VSOP Observations of a Sub-Parsec
Accretion Disk
D.L. Jones, A.E. Wehrle, B.G. Piner & D.L. Meier
Jet Propulsion Laboratory, California Institute of Technology, USA
Abstract
The physical conditions in the inner parsec of accretion disks be-
lieved to orbit the central black holes in active galactic nuclei can
be probed by imaging the absorption of background radio emis-
sion by ionized gas in the disk. We report high angular reso-
lution observations of the nearby galaxy NGC 4261 which show
evidence for free-free absorption by a thin, nearly edge-on disk at
several frequencies. The angular width, and probably the depth,
of the absorption appears to increase with decreasing frequency,
as expected. Because free-free absorption is much larger at lower
frequencies, the longest possible baselines are needed to provide
adequate angular resolution; observing at higher frequencies to
improve resolution will not help.
1 Introduction
The nearby FR-I radio galaxy NGC 4261 (3C270) is a good candidate
for the detection of free-free absorption by ionized gas in an inner ac-
cretion disk. The galaxy is known to contain a central black hole with a
mass of 5× 108 M⊙, a nearly edge-on nuclear disk of gas and dust with a
diameter of ≈ 100 pc, and a large-scale symmetric radio structure which
implies that the radio axis is close to the plane of the sky. At an as-
sumed distance of 40 Mpc, 1 milliarcsecond (mas) corresponds to 0.2 pc.
Previous VLBA observations of this galaxy revealed a parsec-scale radio
jet and counterjet aligned with the kpc-scale jet (see Figure 1). The
opening angle of the jets is less than 20◦ during the first 0.2 pc and < 5◦
during the first 0.8 pc. At 8.4 GHz we found evidence for a narrow gap
in radio brightness at the base of the parsec-scale counterjet, just east
(left) of the brightest peak which we identified as the core based on its
inverted spectrum between 1.6 and 8.4 GHz (see the left part of Figure 2,
from Jones and Wehrle 1997). We tentatively identified this gap as the
signature of free-free absorption by a nearly-edge on inner disk with a
width << 0.1 pc and an average electron density of 103 − 108 cm−3 over
the inner 0.1 pc.
Observations of a Sub-Parsec Accretion Disk
101
Figure 1: VLBA image of NGC 4261 at 8.4 GHz. The contours increase
in steps of √2 starting at ±0.75% of the peak, which is 99 mJy/beam.
The restoring beam is 1.86 × 0.79 mas with major axis PA = −1.3◦.
2 Observations
We observed NGC 4261 at 1.6 and 4.9 GHz with HALCA and a ground
array composed of 7 VLBA antennas plus Shanghai, Kashima, and the
DSN 70-m Tidbinbilla antennas at 1.6 GHz (22 June 1999) and 8 VLBA
antennas plus the phased VLA at 4.9 GHz (27 June 1999). During both
epochs the VLBA antennas at St. Croix and Hancock were unavailable,
as was the North Liberty antenna at 1.6 GHz. Data were recorded
as two 16-MHz bandwidth channels with 2-bit sampling by the Mark-
III/VLBA systems and correlated at the VLBA processor in Socorro.
Both channels were sensitive to left circular polarization.
Fringe-fitting was carried out in AIPS after applying a priori am-
plitude calibration. For VLBA antennas we used continuously measured
system temperatures, for the VLA we used measured TA/TSYS values
with an assumed source flux density of 5 Jy, and for the remaining an-
tennas we used typical gain and system temperature values obtained
from the VSOP web site. Fringes were found to all antennas at 1.6 GHz
except HALCA, but the signal/noise ratio to Shanghai and Kashima was
very low and these data were not used. The a priori amplitude calibra-
tion for Tidbinbilla was dramatically incorrect for unknown reasons. We
calibrated Tidbinbilla by imaging the compact structure of the source
using VLBA data, then holding the VLBA antenna gains fixed and al-
lowing the Tidbinbilla gain to vary. This produced a good match in
Observations of a Sub-Parsec Accretion Disk
102
correlated flux density where the projected VLBA and Tidbinbilla base-
lines overlap. At 4.9 GHz fringes were found to all antennas, including
HALCA. A similar correction to the a priori amplitude calibration for
HALCA and the phased VLA was applied.
In both observations we found that averaging in frequency over both
16-MHz channels in AIPS produced large, baseline-dependent amplitude
reductions even though the post-fringe-fit visibility phases were flat and
continuous between channels. Averaging over frequency within each 16-
MHz band separately fixed this problem. Difmap was used for detailed
data editing, self-calibration, and image deconvolution. Both 16-MHz
bands were combined during imaging.
Imaging within Difmap used uniform weighting with the weight of
HALCA data increased by a factor of 500. Several iterations of phase-
only self calibration, followed by amplitude self calibration iterations
with decreasing time scales, resulted in good fits (χ2 ≈ 1) between the
source model and the data.
3 Results
3.1
1.6 GHz
Although our image at 1.6 GHz does not include data from HALCA,
it does have more than twice the angular resolution of our previous 1.6
GHz image (Jones and Wehrle 1997) due to the addition of Tidbinbilla.
The previous image showed a symmetric structure, with the jet and
counterjet extending west and east from the core. No evidence for ab-
sorption is seen in this image. However, with higher resolution we do
detect a narrow gap in emission just east of the core, at the base of the
counterjet. The width of the gap is less than 2 mas.
3.2
4.9 GHz
We detected fringes to HALCA at 4.9 GHz when the projected Earth-
space baselines were less than one Earth diameter. The HALCA data
fills in the (u,v) coverage hole between continental VLBA baselines and
those to Mauna Kea, and also increases the north-south resolution by a
factor of two. Our 4.9 GHz image is shown in Figure 2. Note that the gap
in emission is again seen just east of the peak. A careful comparison of
brightness along the radio axis at 4.9 and 8.4 GHz shows that the gap is
both deeper and wider at 4.9 GHz, as expected from free-free absorption.
The region of the gap has a very inverted spectrum, the brightest peak
Observations of a Sub-Parsec Accretion Disk
103
Figure 2: Grey-scale images of the nucleus of NGC 4261 at 8.4 GHz
(left) and 4.9 GHz (right), with identical fields of view.
(core) has a slightly less inverted spectrum, and the distant parts of both
the jet and counterjet have steep spectra.
4 Summary
Our observations at 1.6 and 4.9 GHz appear to confirm the free-free
absorption explanation for the sub-parsec radio morphology in NGC
4261. Measurements of the optical depth in the absorbed region and the
distance between the absorption and the core as a function of frequency
will allow the radial distribution of electron density in the inner parsec
of the disk to be determined.
Acknowledgements. We gratefully acknowledge the VSOP Pro-
ject, which is led by the Japanese Institute of Space and Astronautical
Science in cooperation with many organizations and radio telescopes
around the world. This research was carried out at the Jet Propulsion
Laboratory, California Institute of Technology, under contract with the
U.S. National Aeronautics and Space Administration.
References
Jones, D.L. & Wehrle, A.E. 1997, ApJ , 484, 186
|
astro-ph/0510545 | 1 | 0510 | 2005-10-19T01:23:09 | Variable Stars in the Large Magellanic Cloud: Discovery of Extragalactic W UMa Binaries | [
"astro-ph"
] | We observed a field in the disk of the LMC on two consecutive nights in search of rapid variable stars. We have found two pulsating stars of type RRab and delta Sct, and four binary stars, among the latter one sdB or CV below the LMC blue Main Sequence and three very close binary systems on the MS. At least one of the MS binaries, and possibly all three, are the first solar-type (W UMa-type) contact binaries to be detected in any extragalactic system and observed to obey the same Mv = Mv(log P, B-V) calibration as the Galactic systems. Given the selection effects due to small amplitudes at faint magnitudes, the frequency of such binaries in the disk of the LMC with its large spread in population ages is not inconsistent with that in the disk of our Galaxy, and contrasts with the lack of binaries found in earlier observations of the much younger LMC cluster LW55. | astro-ph | astro-ph | Variable Stars in the Large Magellanic Cloud: Discovery of Extragalactic
W UMa Binaries1
Janusz Kaluzny
Copernicus Astronomical Center, Bartycka 18, 00-716 Warsaw, Poland
[email protected]
and
Stefan Mochnacki, Slavek M. Rucinski
David Dunlap Observatory
Department of Astronomy and Astrophysics, University of Toronto
P.O.Box 360, Richmond Hill, Ontario, Canada L4C 4Y6
(mochnacki,rucinski)@astro.utoronto.ca
ABSTRACT
We observed a field in the disk of the LMC on two consecutive nights in search of
rapid variable stars. We have found two pulsating stars of type RRab and δ Sct, and
four binary stars, among the latter one sdB or CV below the LMC blue Main Sequence
and three very close binary systems on the MS. At least one of the MS binaries, and
possibly all three, are the first solar-type (W UMa-type) contact binaries to be detected
in any extragalactic system and observed to obey the same MV = MV (log P, B − V )
calibration as the Galactic systems. Given the selection effects due to small amplitudes
at faint magnitudes, the frequency of such binaries in the disk of the LMC with its large
spread in population ages is not inconsistent with that in the disk of our Galaxy, and
contrasts with the lack of binaries found in earlier observations of the much younger
LMC cluster LW55.
5
0
0
2
t
c
O
9
1
1
v
5
4
5
0
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Subject headings:
clusters -- Magellanic Clouds -- stars: binary -- techniques: photometric
individual: (Large Magellanic Cloud) -- -- galaxies: star
galaxies:
1Based on the data obtained at Las Campanas Observatory, operated by the Carnegie Institution of Washington,
during the University of Toronto time allocation.
-- 2 --
1.
INTRODUCTION
Very little is known about short time scale (< 1 day) variability of stars in the Magellanic
Clouds. Only recently, the availability of large telescopes located at excellent sites has made it
possible to consider time-domain monitoring of the stellar population at brightness levels reaching
and beyond the levels of the Turn-Off Point of the oldest stellar population in LMC at V ≃ 20.5.
The current study addresses the detection and characterization of contact binary stars in a
typical LMC field 1.7 degrees from the center of the LMC. This field was selected taking guidance
from an HST study of the stellar population in the LMC (Smecker-Hane et al. 2002), where it was
called Disk-1. Availability of the archival HST images was one of the reasons for this study as
it permitted us to check for stellar blends and assure better consistency of our results. The field
is characterized by a constant star formation rate from the advanced age of 7.5 to 15 Gyr until
recently. In a companion study on short time-scale variability in the LMC (Kaluzny & Rucinski
2003), the studied field was dominated by a population of the open cluster LW55 with an age
of about 1.5 Gyr, in the presence of an underlying older population with an age about 4 Gyr or
more. We did not find any short-period binaries in or around LW55, but only several short-period
pulsating stars instead.
We present the observations and show the color -- magnitude diagram for stars in the field in
Section 2. The results of the search for short-period variable stars are given in Section 3. Section 4
concludes the paper and summarizes the results.
2. Observations
2.1.
Instruments and observing conditions
We used the Magellan/Baade 6.5m telescope with the TEK5 CCD 2K × 2K camera which had
a focal scale of 0.069 arcsec/pixel. The field of view was 137 × 137 arcsec square. The images were
binned by 2 pixels in both directions before extraction of photometry, with the resulting scale of
0.138 arcsec/pixel. The median seeing during our run was 0.94 arcsec in the V filter and 1.09 arcsec
in the B filter. The binned images significantly oversampled the Point Spread Function (PSF) even
in the cases of the best seeing. The search for variability was done mainly with the V filter (47
images) but also with the B filter (12 images). All V -filter exposures were 600 s while all B-filter
ones were 900 s.
The observations were made on two nights, 2002 January 4/5 and 5/6. We conducted 6.2
hours of variability monitoring on the first night and 7.3 hours on the second night. The nights can
be characterized as gray time, with the fraction of the illuminated area of the Moon disc of 61%
and 50%, respectively.
The image processing was identical to that described in Kaluzny & Rucinski (2003). The initial
-- 3 --
processing of the images was done with standard procedures from IRAF2 with a combination of
the dome and sky flats used for the CCD response flat-field corrections.
The astrometric calibration was based on 48 reference stars from the USNO-B catalogue (Monet
et al. 2003). The random errors for the calibration do not exceed 0.5 arcsec, based on the recovered
coordinates of the USNO-B stars.
2.2. The field Disk-1
Figure 1 shows the sky image in the pixel space (after the 2 × 2 binning), with pixels 0.138
arcsec in size and with a total field of view of 2.3 × 2.3 arcmin square. The field in Figure 1 is
oriented with East to left and North down. The X-coordinate runs W to E, while the Y -coordinate
runs N to S.
2.3. Photometric calibration
The photometric calibrations were done using three of Stetson's standard star fields (Stetson
2000): 2 stars in NGC 1866, 6 stars in E4 -- 108 and 15 stars in NGC 2682 (M67). The fields were
observed with air masses ranging from 1.13 to 1.77. The extinction coefficients as well as the color
terms and zero points of the transformations from the instrumental to the standard BV system
were determined from observations of all 23 standards stars. The adopted formulae were:
v = V − 1.050 + 0.021 × (B − V ) + 0.089 × (X − 1.25)
b = B − 0.702 − 0.142 × (B − V ) + 0.147 × (X − 1.25)
X is an air-mass and lower-case symbols denote instrumental magnitudes derived with the aperture
photometry with the DAOPHOT program (Stetson 1987). Figure 2 shows transformation residuals
for the standard stars.
Table 1 contains the results of our photometry for 4413 stars in the same pixel coordinate
system of as shown in Figure 1. The reference frames in V and B were obtained by collation of,
respectively, 10 and 7 individual frames with the best seeing, in the way as described in detail in
Mochejska et al. (2002). The seeing in the reference images had FWHM of 0.62 arcsec in V and
0.74 arcsec in B. The quoted errors in Table 1 do not reflect stellar variability but are just the
internal errors of the PSF photometry on the reference images.
2IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of
Universities for Research in Astronomy, Inc., under cooperative agreement with the NSF
-- 4 --
Fig. 1. -- The chart of the field in the pixel coordinates, the same as in Table 1. The variable stars
can be found using the boxes marked in the figure, the X, Y coordinates given in Table 2 and the
individual charts in Figure 8.
-- 5 --
Fig. 2. -- Residuals for our photometry of the standard stars, in the sense 'our minus standard"
for V and B magnitudes.
-- 6 --
2.4. Photometric uncertainties
It is difficult to assess reliably errors of single-frame profile photometry in a crowded field such
as Disk-1. Some formal errors, such as those returned by the DAOPHOT software, are often too
optimistic as they tend to miss problems related to blending.
We were not able to compare directly our results with those of Smecker-Hane et al. (2002)
which were discussed only in graphical and verbal form. However, we have analyzed the archival
HST -- WFPC2 data: V : 2 × 500s, F555W, #u4b10905r and #u4b10906r; I: 2 × 300s, F814W,
#u4b10902r and #u4b10904r; the pairs of images were used to remove cosmic rays. We esti-
mated deviations between our photometry and the HST photometry, the latter obtained using the
"HSTphot" software package (Dolphin 2000a,b). While no obvious trends exist in the V magni-
tude differences for 963 common stars in the range 18 < V < 25, there does exist an offset of
∆V = −0.08 ± 0.02 (our results brighter). This offset has not been applied to the photometric
results listed in Tables 1 and the variable stars (Section 3). We note that systematic uncertainties
in the WFPC2 photometry are estimated at about 0.05 mag., and may even reach 0.1 mag. (Piotto
et al. 2002).
To estimate random uncertainties, we compared our results with the HST results for individual
stars. The deviations for the sample of 963 common stars are shown in Figure 3. After accounting
for the systematic shift of −0.08, binning the deviations in one magnitude intervals, and with
an assumption that errors from both sources add quadratically, we estimate the rms scatter at
σV ≃ 0.02 − 0.04 for for V < 21; it increases to 0.09 at V = 22 and to 0.11 at V = 23.
Much more reliable than the above estimates are estimates of errors in the search for variable
stars, obtained by comparison a large number of images analyzed using the image differencing
technique. They are discussed in Section 3 in a discussion of amplitudes of detectable variable
stars.
2.5. The color magnitude diagram
The photometric results for the whole field are plotted on the color -- magnitude diagram (CMD)
in Figure 4. As discussed by Smecker-Hane et al. (2002), the Disk-1 field contains populations of
different age with clear indications that the star formation rate in this part of the LMC was constant
over a long time. We can see the red horizontal branch (at about V ≃ 19.5) and the Main Sequence
turn-off point (at about V ≃ 21) of the old population as well as a sequence of young stars extending
to V ≃ 18.
-- 7 --
Fig. 3. -- ∆V differences between our photometry and HST photometry (after allowance for the
offset of −0.08 mag) can be taken as a measure of random errors at different magnitude levels.
Estimates of the rms errors at one magnitude intervals are shown by vertical bars.
-- 8 --
Fig. 4. -- The CMD for the Disk-1 field. The six variable stars are marked with large filled circles
at their median magnitudes and color indices. The vertical bar for V4 shows its observed ranges
of variability in V ; for the remaining variables the ranges are smaller than the symbol size. The
absolute magnitude scale on the right vertical axis is based on the assumption of (m − M )0 = 18.5
and EB−V = 0.10.
-- 9 --
3. Variable stars in Disk-1
3.1. Techniques and errors
A search for potential variable stars in the field Disk-1 was performed with the image sub-
traction package ISIS V2.1 (Alard & Lupton 1998; Alard 2000). Two methods were used to detect
potential variable objects. First, we applied procedures which are included in the ISIS package and
which are based on analysis of residual images. The second method relies on extraction -- still within
the ISIS package -- of light curves for all stellar objects whose positions had been determined on
template images with DAOPHOT/Allstar (Stetson 1987). Extracted light curves are subsequently
examined for the presence of any possible periodic variations with a suite of programs based on the
"AoV" algorithm (Schwarzenberg-Czerny 1997).
Figure 5 shows the rms versus the magnitude diagram for 6243 stars whose light curves were
analyzed (this number is by about one half larger than the number of stars that went into the
CMD). The rms errors were derived after rejection of the two most extreme maximal and minimal
data points from data for each star. Not all points with large σ in the figure correspond to
physical variables as automatic processing led to inclusion of a few easily identifiable stars with
poor photometry, mostly in the extended wings of over-exposed bright stars. The figure suggests
that assuming a 5 × σ detection threshold, we should be able to detect variable stars with peak-to-
peak amplitudes of ∆V ≃ 0.05 at V = 21.5 and ∆V ≃ 0.10 at V = 22.5. It is encouraging to note
that the data were taken during gray time. Clearly, even during gray time, one may use the Baade
telescope to look for variable stars in the LMC beyond the turnoff region at V ≃ 22.
3.2. Results of the variable star search
We have detected six variable stars in the Disk-1 field. They are marked on the color-magnitude
diagram of the field in Figure 4 and are listed in Table 2, where their CMD numbers, X, Y and
equatorial J2000 coordinates, the maximum and minimum V and median values of (B − V ) are
given. The listed values of (B − V ) are uncertain because of the non-simultaneous nature of our
B and V observations and because the exposure times were rather long (15 minutes in B). The
B − V data given in Table 2 are slightly different than the ones in Table 1 due to stellar variability
along with differences in how the mean photometric values were determined.
The light curves of the variables3 are shown in Figures 6 and 7, while the finding charts in
pixel coordinates of Figure 1 are shown in Figure 8. Figures 6 and 7 give the V and B − V curves,
3The
tables
of
individual V
and B magnitudes
and interpolated B − V
are
available
from:
http://astro.utoronto.ca/rucinski/LMC-Disk1/Var* Disk1.dat
-- 10 --
Table 1. Photometric data for LMC Disk-1
#
1
2
3
4
5
6
7
8
9
10
11
12
X
Y
RA[deg]
Dec[deg]
V
σV
(B − V )
σ(B − V )
16.43
16.95
17.06
17.85
17.86
18.43
18.73
18.87
19.23
19.99
20.07
20.12
841.77
736.51
175.30
548.81
423.94
1002.93
760.67
461.99
88.02
662.28
56.83
488.31
77.92661 −71.19987
77.92596 −71.19575
77.92218 −71.17359
77.92495 −71.18848
77.92420 −71.18362
77.92822 −71.20638
77.92633 −71.19669
77.92456 −71.18511
77.92155 −71.16988
77.92588 −71.19287
77.92130 −71.16853
77.92488 −71.18614
22.409
22.363
22.357
17.787
22.716
21.848
22.535
21.001
23.817
23.804
21.416
22.957
0.019
0.008
0.006
0.004
0.008
0.035
0.007
0.004
0.015
0.021
0.007
0.008
0.580
0.339
0.350
0.831
0.498
0.399
0.429
0.172
0.533
0.494
0.389
0.435
0.031
0.016
0.012
0.008
0.017
0.056
0.013
0.008
0.037
0.041
0.016
0.016
Note. -- Table 1 is presented in its entirety in the electronic edition of the Astronomical
Journal. A portion is shown here for guidance regarding its form and content. The stars
positions are given in the pixel X and Y coordinates of the reference image, as in Figure 1,
and in the equatorial coordinates in the J2000 system. The formal errors given here can be
used to compare relative uncertainties; the external estimates of the errors are discussed in
the text.
Fig. 5. -- The rms errors of photometry based on the inter-comparison of the individual images
used for detection of variability. The two panels show separately photometric errors in V and B
bands versus the respective magnitudes.
-- 11 --
Fig. 6. -- The light and color index curves for variables V1 -- V3. For each star, the left panel shows
the phased data with T0 and the period as given in Table 2, while the two right panels show the
individual nightly data. The vertical scale has the same range of ∆V = 0.3 and ∆(B − V ) for all
stars, but the magnitude and color index levels are different.
-- 12 --
Fig. 7. -- The same as in Figure 6, but for the variables V4 -- V6. Note that for V4, the magnitude
scale is expanded 3 times relative to that for the remaining stars.
-- 13 --
Fig. 8. -- The finding charts of the variable stars showing 40 × 40 pixels of our reference image
(Figure 1) centered on each of the variable stars.
-- 14 --
the latter for the interpolated moments of the B observations. We do not discuss the B magnitude
light curves in this paper because of their low precision.
We inspected the archival WFPC2 images (particularly the image WFPC2ASN U4B10905B)
and found that all our detected variables are free from blending. We also found that our RA/Dec
coordinates of the variables agree with the WCS of the WFPC2 image to 0.6 arcsec.
Obviously, due to the short duration of the program, we could detect only short time scale
variables. On the other hand, we could probe relatively deeply and could detect variables among
stars of very moderate brightness reaching as far down the main sequence as solar stars, in the
LMC at V ≃ 22 − 23. The limited scope of the project is fully confirmed by the characteristics of
the detected variables: As we can see in Fig. 4, four of the stars are located in the Main Sequence
of the LMC; these are V1, V2, V5, V6. One of them, V2, is a δ Scuti variable, while the remaining
ones are short period binary systems. V1, V2 and V6 are located in the turn-off region of the old
population Main Sequence within +0.28 < B − V < +0.39 and V ≃ 21. We also detected one
RR Lyr-type (RRab) variable and one very blue variable below the Main Sequence. We comment
on the individual objects below.
• V1 (#1662) -- A short period eclipsing binary with an amplitude of light variations of ∆V ≃
0.15, period P = 0.49 d and B − V = 0.28. Although the light curve may suggest a close,
but detached binary, MV = 2.0 predicted from (m − M )0 = 18.5 and EB−V = 0.1 perfectly
agrees with the contact-binary calibration of Rucinski & Duerbeck (1997)(see also Rucinski
(2000)), MV (cal) = 2.0.
• V2 (#1984) -- A short-period (0.0675 d) δ Sct or SX Phe pulsating star with ∆V = 0.06 at
V = 21.15. The light curve is relatively well defined thanks to the folding of several periods.
• V3(#1999) -- An interesting blue star well below the Main Sequence at V ≃ 22.45 and B −V ≃
−0.1 (independent photometry on the reference image of Section 2.5 gives B − V ≃ −0.05).
Variability is rapid; if it is a binary, as individual nightly observations suggest, then the period
is 0.2607 day, but the period may be actually 1/2 of this value. We verified that the blue
color of the star is not due to blending, which is possible in such a crowded field as Disk-1
because the archival WFPC2/HST images mentioned above clearly show that V3 is indeed
very blue. The observed luminosity of V3 is consistent with a relatively bright Cataclysmic
Variable in the LMC. The variable is isolated and bright enough that low-resolution spectra
could be obtained from the ground.
• V4 (#2618) -- This is an RRab pulsating star with well defined light variations, but an
incomplete light curve. We assumed that two cycles elapsed between the two nights, P=0.538
d. Since we have not captured the light maximum, we have been unable to determine the
initial (maximum light) epoch T0 for this star; the maximum brightness is possibly above
V = 18.8. Because the light curve is incomplete, we cannot relate photometric properties of
the star to the horizontal branch of the LMC, but the properties are certainly consistent.
-- 15 --
• V5 (#3175) -- A rather typical contact binary well within the Main Sequence of the LMC,
some +2 magnitudes below the Turn-Off Point, at Vmax = 22.57. Assuming B − V = 0.47,
EB−V = 0.1 and P = 0.3108 d, the calibration of Rucinski & Duerbeck (1997) predicts
MV (cal) = 3.5; directly, with (m − M )0 = 18.5, one obtains MV = 3.8. This is consistent
within the current uncertainties. V5 is the first certain W UMa-type binary identified in
other galaxy.
• V6 (#4284) -- A close eclipsing binary with Vmax ≃ 20.8, B − V = 0.39 and the period of
0.469 d (assuming the observations cover two cycles). From the distance modulus, MV = 2.0,
while -- assuming that the binary is really a contact one -- the calibration of Rucinski &
Duerbeck (1997) gives MV (cal) = 2.4; the light curve does not look like that of a contact
binary, however.
4. Summary and conclusions
We have photometrically observed a region of the Large Magellanic Cloud -- called "Disk-1"
by Smecker-Hane et al. (2002) -- on two consecutive nights in search of short time scale variable
stars. The time monitoring led to the discovery of four short-period eclipsing binaries and of two
pulsating stars (RRab and δ Sct). The results clearly show that solar-type stars are accessible for
such monitoring in the LMC down to V ≃ 22.5 − 23.
It appears to be significant that the field Disk-1 contains short period binaries. We discovered
four such binaries, in contrast to the results for the LMC field of LW55 (Kaluzny & Rucinski 2003)
which was observed in a practically identical way and where we detected only very low amplitude,
pulsating variables of δ Scuti, SX Phoenicis or γ Doradus type, and no binaries. While small-number
statistical fluctuation is still a possibility, the reason may be in the uniform distribution of stellar
ages in the Disk-1 field (Smecker-Hane et al. 2002), contrasted with the population dominated by
LW55 itself with an age of 1.5 Gyr.
Obviously, the number of short-period binaries is still small, 4 among 6243 monitored including
possibly 3 contact binaries, compared with the Galactic field where typically one among 500 FGK
dwarfs is expected to be a contact binary (Rucinski 2002). However, all binaries detected by us have
moderately large amplitudes, ∆V > 0.1 − 0.15, which may result from selection effects operating at
the faint levels of V ≃ 24 in a very crowded field; we simply could not detect low amplitude binaries.
Numbers of close binaries apparently increase rapidly with decreasing amplitude of light variations
(Rucinski 2001). This is confirmed by the complete (∆V > 0.05) sample of bright, short-period
binaries in the solar neighbourhood, mostly from the Hipparcos catalog, with V < 7.5 (Rucinski
2002).
In this sample, only about 1/3 of all binaries have amplitudes ∆V > 0.10 and almost
1/2 of them have amplitudes ∆V < 0.15. Taking this amplitude selection effect into account, the
discrepancy between the observed 3 and the expected 8 contact binaries does not appear to be
significant, assuming that about 2/3 of the monitored stars were FGK dwarfs.
-- 16 --
We thank the reviewer Dr. Wayne Osborn for very careful checking of our paper and for several
useful suggestions.
Research support from the Ministry of Scientific Research and Informational Technology,
Poland to JK (grant 1 P03D 001 28) and from the Natural Sciences and Engineering Council
of Canada to SWM and SMR is acknowledged here with gratitude.
This work is based on observations with the NASA/ESA Hubble Space Telescope, obtained
from the Data Archive at the Space Telescope Science Institute, which is operated by the Associ-
ation of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These
observations are associated with program #7382
REFERENCES
Alard, C. 2000, A&AS, 144, 363
Alard, C., & Lupton, R. H. 1998, ApJ, 503, 325
Dolphin, A. E. 2000a, PASP, 112, 1383
Dolphin, A. E. 2000b, PASP, 112, 1397
Kaluzny, J., & Rucinski, S. M. 2003, AJ, 126, 237
Mochejska, B. J., Stanek, K. Z., Sasselov, D. D., & Szentgyorgyi, A. H. 2002, AJ, 123, 3460
Monet D. et al, 2003, AJ, 125, 984
Piotto, G., King, I. R., Djorgovski, S. G., Sosin, C., Zoccali, M., Saviane, I., De Angeli, F., Riello,
M., Recio Blanco, A., Rich, R. M., Meylan, G. & Renzini, A. 2002, A&A, 391, 945
Rucinski, S. M. 2000, AJ, 120, 319
Rucinski, S. M. 2001, AJ, 122, 1007
Rucinski, S. M 2002, PASP, 114, 1124
Rucinski, S. M., & Duerbeck, H.W. 1997, PASP, 109, 1340
Schwarzenberg-Czerny, A. 1997, ApJ, 489, 941
Smecker-Hane, T. A., Cole, A. A., Gallagher III, J. S., & Stetson, P. B. 2002, ApJ, 566, 239
Stetson, P. B. 1987, PASP, 99, 191
Stetson, P. B. 2000, PASP, 112, 925
This preprint was prepared with the AAS LATEX macros v5.2.
-- 17 --
Table 2. Variable stars in Disk-1
V
#
X
Y
Vmax
Vmin
(B − V ) RA(hh:mm:ss) Dec(dd:mm:ss)
Type
Period(d)
T0
1
2
3
4
5
6
1662
1984
1999
2618
3175
4284
381.82
449.61
453.81
593.36
713.90
971.90
20.84
953.11
21.12
484.19
22.39
866.78
609.87 <18.80
22.57
814.92
630.27
20.79
20.97
21.18
22.57
19.56
22.78
20.91
0.28
0.33
−0.10
0.38
0.47
0.39
5:11:53.10
5:11:54.51
5:11:55.02
5:11:58.71
5:12:02.35
5:12:09.52
−71:12:14.6
−71:11:09.4
−71:12:02.0
−71:11:25.8
−71:11:53.2
−71:11:26.2
EA/EW:
δ Sct
sdB/CV
RRab
EW
EA/EW:
0.491(1):
0.0675(1)
0.2607(2)
0.538:
0.3108(1)
0.469(1):
79.715
79.568
80.67
80.625
80.69
Note. -- The variable star numbers, as used in the text, are given in the first column, while identifications in Table 1 are in the
second column. The full names which conform to the International Astronomical Union recommendations are "Disk1 -- LCO -- V...".
The J2000 equatorial coordinates were obtained through an astrometric frame solution using positions of 48 stars from the USNO-B
catalogue; this solution reproduces the J2000 coordinates of these stars with residuals not exceeding 0.5 arcsec in RA and Dec. Note
that the V and B − V data differ slightly between this table and Table 1 due to different photometric methods used; the differences
can be taken as indication of external uncertainties combined with the genuine variability.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.