Introduction To Stellar Pulsation: 7.1 Some Basic Terminology

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Chapter 7

Introduction to Stellar Pulsation

Throughout this course, we focus on the static structure of a star. More


precisely, we assume that the star is in mechanical equilibrium all the time.
Nonetheless, under suitable conditions, the equilibrium configuration of certain
stars is unstable against perturbation. Such perturbation may lead to stellar
pulsation. Study of stellar pulsation is important for astronomers for the fol-
lowing reasons: 1. Distance measurement. Certain stellar pulsation leads to
pulsation variable. The most famous one is the Cepheid variable, which is
well-known for its period-luminosity relationship that has been used as
standard candle to set the cosmological distance scale. 2. Stellar structure.
Our Sun and other main-sequence stars also show pulsations, albeit with much
smaller amplitudes. By observing different oscillation modes of a star, one can
infer its interior structure. This is known as asteroseismology (or helioseis-
mology for the specific case of the Sun).

7.1 Some basic terminology

The instability strip is a narrow region in the Hertzsprung-Russell diagram


where most of the pulsating variables lie (except the Mira variables), e.g., it
intersects with the horizontal branch at RR Lyrae, crosses the supergiants at
Cepheids, and crosses the white dwarf region at ZZ Ceti.

The pulsations in Cepheids, Miras, and RR Lyrae are radial. They show peri-
odic contraction and expansion with periods from 1 to 100 days. On the other
hand, the Sun, δ Scuti, and white dwarfs are non-radial pulsators. Differ-
ent regions on the stellar surface contract and expand differently. Non-radial

71
CHAPTER 7. STELLAR PULSATION 72

Figure 7.1: Location of pulsating variables on the H-R diagram.

oscillations can be described by the spherical harmonics


v
m t 2l + 1 (l − m)! m
u
Ylm (θ, φ) = P (cos θ)eimφ .
u
(−1) l (7.1)
4π (l + m)!

Stellar pulsation is similar to sound waves in musical instruments. They are


standing waves with the center of the star as a node and the stellar surface as
an antinode. In the fundamental mode (l = 0), there is no other nodes and
the entire star pulse in a single direction at a given time. The first overtone
(l = 1) correspond to one node between the center and the surface, the second

Type Period Mode


Mira 100–1000 d radial
Cepheids 1–50 d radial
W Virginis 10–20 d radial
RR Lyrae 1 hr–1 d radial
δ Scuti 1–3 hr rad. & non-rad.
β Cephei 3–7 hr rad. & non-rad.
ZZ Ceti 100–1000 s non-rad.

Table 7.1: Some common pulsating variables.


CHAPTER 7. STELLAR PULSATION 73

overtone (l = 2) has two, and so on. A star can exhibit several modes at the
same time. Cepheids are believed to oscillate in the fundamental mode, while
both the fundamental mode and the first overtone can exist in RR Lyrae and
Miras.

7.2 A simple model

In this introduction, we will focus on the theory of the simplest possible model
of stellar pulsation, namely, adiabatic radial pulsation in the small amplitude
limit. Further analysis can be found in the following review articles: J. P. Cox,
Reports on Progress in Physics 37, 563 (1974); A. Gautschy and H. Saio, Annual
Review of Astronomy and Astrophysics 33, 75 (1995); A. Gautschy and H. Saio,
Annual Review of Astronomy and Astrophysics 34, 551 (1996).

The basic idea is simple, when a layer is compressed, it heats up and becomes
more opaque to radiation. Radiative diffusion slows down, heat, and hence
pressure, build up. This eventually pushes the layer outward. The expansion
cools down the material, and it becomes more transparent to radiation, such
that heat can escape more easily and the pressure subsequently drops. Finally,
the layer falls back and the cycle starts over.

Here we use the so-called Lagrange approach which focuses on the motion of
individual particles in a star. (An alternative approach is that of Euler, which
focuses on the evolution of thermodynamic variables such as density and pres-
sure. Both approaches give the same physics but the later one is sometimes
more effective in solving macroscopic problems involving fluid.)

The conservation of mass (Eq. 2.2 in Chapter 2) in the time independent case)
should be rewritten as
∂m
= 4πr2 ρ(r, t) . (7.2)
∂r
In contrast, the equation of hydrostatic equilibrium (Eq. 2.3 in Chapter 2)
becomes more complicated in the time dependent case. We assume that the
oscillation is radial so that all physical parameters are functions of r and
t only. In this case, the mass of a small spherical shell from radius r to
r + ∆r equals 4πr2 ρ(r, t)∆r. The gravitational force acting on this shell is
−Gm(r, t)4πr2 ρ(r, t)∆r/r2 . The (gas plus radiation) pressure gradient acting
on this shell is ∂P/∂r. Hence, by Newton’s second law,

G4πr2 ρm ∂P
4πr2 ρr̈ = − 2
− 4πr2 . (7.3)
r ∂r
CHAPTER 7. STELLAR PULSATION 74

In other words,
Gm ∂P
r̈ = − 2
− 4πr2 . (7.4)
r ∂m

Assuming that the oscillation is adiabatic and the star is made up of an ideal
gas, pressure is given by
P = Kργa (7.5)
for some constant K.

Analytical solution to these three partial differential equations, namely, Eqs. (7.2),
(7.4) and (7.5) are difficult to find. Nonetheless, in most cases, the amplitude of
oscillation of a star about its equilibrium position is small enough. In this case,
we need only to keep up to the first order term in the Taylor series expansion
about the equilibrium configuration of the star. Besides, the oscillation of the
star can be well approximated by a S.H.M. In other words, using m and t are
the two independent variables, we may write

P (m, t) = P0 (m)[1 + p(m)eiωt ] , (7.6)

r(m, t) = r0 (m)[1 + x(m)eiωt ] , (7.7)


and
ρ(m, t) = ρ0 (m)[1 + d(m)eiωt ] , (7.8)
where P0 (m), r0 (m) and ρ0 (m) are the equilibrium pressure, radius and density
on the sphere whose mass contained equals m.

Substituting the above equations into the three coupled partial differential equa-
tions and keeping up to the first order terms, we obtain
P0 ∂p Gm(p + 4x)
= ω 2 r0 x + , (7.9)
ρ0 ∂r0 r02
∂x
r0 = −3x − d , (7.10)
∂r0
and
∂x
p = −3γa x − γa r0 . (7.11)
∂r0
Eliminating p and d, we have
! " #
∂ ∂x 4 ∂ Gmρ0 γa ∂x ρ0 Gm
γa + (γa x) − + (4 − 3γa ) + ω 2 x = 0 .
∂r0 ∂r0 r0 ∂r0 P0 r0 ∂r0 P0 r03
2

(7.12)

So the problem of stellar oscillation becomes the problem of finding eigenvalues


ω 2 for above linear partial differential equation. If the eigenvalue ω 2 > 0, the
CHAPTER 7. STELLAR PULSATION 75

solution is oscillatory which corresponds to an radial oscillation of the star. If


ω 2 < 0, the solution contains an exponentially decaying or growing term which
indicates instability.

Eq. (7.12) is still too complicated to be solved analytically in general. Never-


theless, we have one specific meaningful case in which this eigenvalue problem
has been investigated in detail. This is the case of (1) the adiabatic expansion
coefficient γa is a constant throughout the star; and (2) x is m independent. In
this case, Eq. (7.12) becomes
Gm
ω 2 = (3γa − 4)
r03
4πGρ̄
= (3γa − 4) , (7.13)
3
where ρ̄ is the mean density of the star. In other words, the oscillation of the
star is stable if γa > 4/3; and the period of oscillation Π in this case is given by
" #1/2
3
Π = 2π . (7.14)
4π(3γa − 4)Gρ̄
For non-relativistic ideal gas, γa = 5/3. Put this into the equation above, we
have:

Type Period
RR Lyrae ≈ 0.5 d
Classical Cepheids ≈ 7d
W Virinis ≈ 15 d

This agrees reasonably well with observations.

We have not finished our discussions yet. In reality, our linear analysis in the
adiabatic regime is an overly simplified approximation. Actually, if we include
higher order effects, radial oscillation of a star will be damped out quickly unless
there is an effective way to pump energy into the oscillator. Indeed, pulsating
variables, such as Cepheids, occupy only a restricted strip on the H-R diagram,
strongly suggesting that an effective excitation mechanism exists only under
rather restrictive conditions.

7.3 κ-mechanism and the instability strip

Detail analysis of the excitation mechanism is rather involved. The most famous
one is the so-called κ-mechanism. This mechanism assumes that the opacity
CHAPTER 7. STELLAR PULSATION 76

increases upon compression in some region of the star so that the radiative
luminosity is blocked in the compression phase of oscillation. Therefore, the
region gains thermal energy in compression phase and loses thermal energy in
expansion phase.

Recall that for most part of an “ordinary” star, κ is well approximated by


κ0 ρT −7/2 (Kramer’s law). Hence, upon adiabatic compression, the opacity ac-
tually decreases. Thus, the condition for κ-mechanism is not satisfied in most
stars. Nevertheless, such a condition holds in regions where there is partial
ionization of H and Hei and/or Heii. In fact, κ-mechanism due to such par-
tial ionization zone is responsible for the excitation of pulsation in a number
of pulsating variables including RR Lyrae, classical Cepheids and W Virginis.
Because the condition of onset of the κ-mechanism due to partial ionization is
very stringent, pulsating variables are rare and can only be found in restricted
regions on the H-R diagram.

We will not go into the detail mathematics here, but only give a qualitative
description of the κ-mechanism and its connection to the instability strip. In
the partial ionization regions, part of the energy input from compression can
be turned into more ionization instead of heating. Therefore, it is possible to
increase in density more than in temperature. Hence, κ can be increased due
to compression. Later in the cycle, electron-ion recombination during decom-
pression can release energy to lower the opacity.

Where is the κ-mechanism expected to occur? Location of the partial ionization


zone in a star depends strongly on the effective temperature. If the temperature
is too high, the ionization will be close to the surface, resulting in insufficient
mass to drive sustainable oscillations. On the other hand, if the temperature is
too low convection in the outer layer will strongly suppress the κ-mechanism.
Therefore, we expect a narrow, near vertical strip in the H-R diagram.

7.4 ε-mechanism

For completeness, we also mention the ε-mechanism. It was proposed before


the κ-mechanism to explain stellar pulsations. It suggests that compression at
the center of a star can lead to higher density and temperature, thus, increases
the thermonuclear energy production rate. However, detailed calculations show
that the oscillation amplitude at the central regions of most stars are small, such
that this effect is insignificant. However, this is still an important mechanism
in massive stars.
CHAPTER 7. STELLAR PULSATION 77

7.5 Cepheid variables

Classical Cepheids (or Cepheids for short) are luminous supergiants. It is named
after the first known example, δ-Cephei, which is still an important calibrator
for the period-luminosity relation (see below). Their immediate precursors are
massive young O or B type stars. Their periods are about 2 to 60 d, mean
luminosities about 300 to 40000 L , mean surface temperature about 4000 to
8000 K, and peak-to-peak magnitude variation of about 0.4 to 1.4 mag. Its
stellar atmosphere is cold enough to have a layer of partially ionized Hei and
Heii. Hence, κ-mechanism leads to stellar pulsation. (See Figure 7.2 below
for a typical light curve, color change and radial velocity of a classical Cepheid
variable.)
mv

3.5

4.0

4.5
0.0 0.5 1.0
Teff
6500

6000

5500
Vr (kms −1) 0.0 0.5 1.0
20

10
0
−10
−20
0.0 0.5 1.0
Phase

Figure 7.2: Location of δ Cephei in the sky, and the variation of visual magni-
tude, effective temperature and radial velocity.

In 1912, Henrietta Leavitt accidentally found that the period and luminosity
of classical Cepheid variables are related. In a more modern language, such a
period-luminosity relation is found to be
!
L
log = 2.43 + 1.18 log Π . (7.15)
L

That is, the longer the period, the brighter the star is. This famous relation
is also called the Leavitt Law after her. Note that the above relation is valid
plus or minus a few percent error.
CHAPTER 7. STELLAR PULSATION 78

We may understand the period-luminosity relation as follows. First, the peak-


to-peak magnitude difference is relatively small, hence we may assume that
the effective surface temperature of a classical Cepheid to be a constant. The
luminosity of the star L ≈ 4πσR2 T 4 is therefore a function of stellar radius R
only. Although classical Cepheids are supergaints, the homology relation for
main sequence star still holds approximately. Hence, we have an approximate
mass-luminosity relation L ∼ M α for classical Cepheids. Finally, we have the
period-mean density relation Π ∼ ρ̄−1/2 . Combining these three constraints,
we arrive at a period-luminosity relation for classical Cepheid variables L ∼
Π4α/(3α−2) . For α ≈ 3, we obtain L ∼ Π1.7 . This is about the same order as the
empirical formula above, but we cannot derive the exact value since the mean
density approximation is oversimplified.

Indeed, the above argument is rather general and can be applied to quite a
number of pulsation variables including W Virginis and Mira type variables.
A notable exception is the RR Lyrae variables. Observations suggest that the
absolute magnitudes of RR Lyrae variables fall into a narrow range of 0.6 ± 0.2.
Hence, within about 10% accuracy, we may assume that all RR Lyrae stars are
of the same luminosities and hence can be regarded as standard candles.

7.6 Cepheid variables in the cosmic distance


ladder

Since the period is much easier to measure to high precision, as compared to


other observables such as spectrum or position. The period-luminosity relation
immediately turned Cepheids into standard candles, offering a very powerful
tool for distance measurements. Edwin Hubble observed Cephieds in other
galaxies to discover Hubble’s law. Even today, Cepheids play an important
role in the construction and calibration of the cosmic distance ladder, up to the
order of 10 Mpc (see the Hubble Space Telescope Key Project).
Chapter 8

Supernova Remnants

The aftermath of a supernova explosion is a cloud of hot gas expanding into


the surrounding interstellar medium (ISM), known as a supernova remnant
(SNR). SNRs play crucial roles in heating and chemical evolution of galaxies.
The shock waves also trigger of stellar formation, thus closing the cycle of stellar
evolution. SNRs also allow us to understand how heavy elements are synthesized
and populated in the Universe.

In this chapter, we review the basics of shock wave and evolution of SNRs. Al-
though we only talked about core collapse supernovae in this course, everything
we discuss in this chapter also applies to type Ia supernovae.

8.1 Shock waves

Figure 8.1: The schematic of a shock. (a) The velocity of the shock front is U
and the upstream is stationary. (b) The reference frame where the shock is at
rest. v1 and v2 in the equations are defined in this frame.

The ejecta materials from a supernova is highly supersonic. When encountering

79
CHAPTER 8. SUPERNOVA REMNANTS 80

the surrounding ISM, the flow is forced to slow down to subsonic speed. This
leads to discontinuity of the flow velocity, which is called a shock.

A shock is an irreversible wave that causes a transition from supersonic to


subsonic flow. Because the shock thickness is of the order of the mean free
path of the gas particles, it is usually much smaller than the length scales of
gradients in the gas on either side of the shock, In the following discussion,
we regard it as infinitely thin. A schematic illustration of a shock is shown in
Figure 8.1. The flow is divided into upstream (denoted as 1) and downstream
(denoted as 2) regions by the shock front. It is easier to do the calculation in the
shock reference frame (right panel). The flow is characterized by the pressure,
temperature, density, and flow speed, and we can write down three equations:
conservation of mass
ρ1 v1 = ρ2 v2 , (8.1)
conservation of momentum

P1 + ρ1 v12 = P2 + ρ2 v22 , (8.2)

and conservation of energy


p1 1 2 p2 1 2
e1 + + v1 = e2 + + v , (8.3)
ρ1 2 ρ2 2 2

where e1/2 is the specific energy given by e = P/(γ − 1)ρ and γ is the adiabatic
index. Combining with the ideal gas law, one can show that

ρ2 2(M2 − 1)
=1+ , (8.4)
ρ1 2 + (γ − 1)M2

p2 2γ(M2 − 1)
=1+ , (8.5)
p1 γ+1
T2 2(γ − 1)(γM2 + 1)(M2 − 1)
=1+ , (8.6)
T1 (γ + 1)M2
q
where M = v1 /c1 is the Mach number upstream (c1 = γp1 /ρ1 is the sound
speed upstream). In strong stock / highly supersonic shock regime, i.e. M  1,
the ratio of the velocities becomes
v1 ρ2 γ+1
= ≈ . (8.7)
v2 ρ1 γ−1

For monatomic particles, γ = 5/3, this gives v1 /v2 = 4. This relation is impor-
tant when we discuss diffusive shock acceleration in Chapter 10. In addition,
Eq. 8.6 shows that the shocked gas could have very high temperature in strong
shocks.
CHAPTER 8. SUPERNOVA REMNANTS 81

8.2 Evolution of supernova remnants

The enormous explosion energy of a supernova creates a cavity in the ISM


known as an SNR, which gradually expands and the diameter can eventually
reach up to several hundred parsecs. The time evolution of an SNR can be
roughly divided into four phases:

1. Free expansion phase


A supernova typical has kinetic energy of ESN ∼ 1051 erg. This is converted into
the ejecta velocity vej as
1
ESN ∼ Mej vej2 (8.8)
2
!1/2 !−1/2
4 −1 ESN Mej
vej ∼ 10 km s . (8.9)
1051 erg M
This suggests an initial expansion velocity of a few percent of speed of light.
The first stage is the free expansion phase, during which the stellar ejecta
retain their initial velocity and expand linearly. Therefore, the blast wave radius
increases with time t as vej t. The blast wave starts to slow down only after it
swept up comparable mass as the ejecta mass. The remnant then reaches the
end of this first phase. The swept-up mass is related to the SNR radius r by
Mswept−up ∼ (4π/3)r3 ρISM . Hence, at the end of this phase, the remnant reaches
a radius !1/3 !−1/3
Mej ρISM
r1 ∼ 2 pc . (8.10)
M 10−24 g cm−3
Dividing this by the velocity gives an age estimate when this occurs,
!−1/2 !5/6 !−1/3
ESN Mej ρISM
t1 ∼ 200 yr . (8.11)
1051 erg M 10−24 g cm−3

2. Sedov phase
In the second phase, the energy loss through radiation is still negligible. We
can then assume that ESN = constant. Also, we now have Mswept−up > Mej ,
the total energy can be approximated by ESN ∼ 21 Mswept−up v 2 . Rewrite the
equation as
ESN ∝ ρISM r3 ṙ2 , (8.12)
which can be integrated to yield
r ∝ (ESN /ρISM )1/5 t2/5 . (8.13)
What we derived above is also called the Sedov-Taylor self-similar solution to
a spherical shock, which was first derived to estimate the energy of an atomic
CHAPTER 8. SUPERNOVA REMNANTS 82

bomb based on dimensional analysis. This phase is also called the Sedov phase.
Unlike in the first phase, the expanding wave now decelerates as it expands.
Clearly, material further out decelerates first, so that material further in starts
to run into the outer shells. This heats up the outer shell and can produce a
double shock structure with complex flow patterns.

The temperature in the shocked gas is high. In the limit of strong shocks, the
ratio of the temperatures behind and in front of the shock is given by Eq. 8.6

T2 2γ(γ − 1)M2
≈ for M  1. (8.14)
T1 (γ + 1)2

In the case of SNR, the Mach number is large and the temperature can reach the
order of 106 K. Such hot plasma strongly emits X-rays, mainly through thermal
bremsstrahlung.

3. Snowplow phase
When temperatures fall to less than a million K, some ions start to recombine
to form atoms, resulting in line emission and strong cooling of the gas. When
radiative losses start to affect the kinetic energy, the Sedov phase ends. This
occurs at times of about 10 kyr. The shell still propagates with constant radial
momentum and piles up the ambient material like a snowplow. By conservation
of momentum,
d d 4π
  
3
(M v) = ρr ṙ = 0. (8.15)
dt dt 3
Integrating this with respect to time yields
" #1/4
4v0 (t − t0 )
r = r0 1+ , (8.16)
r0

which, at late times, has the solution r ∝ t1/4 . At this stage, the gas has
temperatures of around 104 K and radiate strong line emission, and radiates
strongly in the optical waveband. This is called the snowplow or radiative
phase.

4. Final phase
In the final stage of the SNR, the stellar debris merges with the ISM. This
happens when the speed of the ejecta become comparable to the sound speed
of the ISM, which usually takes about 100 kyr after the supernova explosion.
CHAPTER 8. SUPERNOVA REMNANTS 83

Figure 8.2: An illustration of the four different phases in the dynamics of a


supernova remnant (SNR). Rough estimates for the temperatures and velocities
at the end of each phase are given.

You might also like