Sie sind auf Seite 1von 15

Basic concepts in stellar physics

Anthony Brown
brown@strw.leidenuniv.nl
10.02.2010

Abstract. These notes provide extra material on the basics of stellar physics. The aim is to provide
an overview of the nature of stars using only elementary physics. Some concepts that are fundamental to stellar physics are introduced and the order of magnitude of astrophysical quantities is
fixed through simple estimates. Most subjects treated in this note will be discussed in more detail
using the chapters on stellar structure and evolution from Ostlie & Carroll (2007). The material
below is extracted from the first chapter in Phillips (1998).

Revision History
Issue
3
2
1
0

Rev. No.
0
0
0
0

Date
09.02.2009
09.02.2009
26.02.2008
24.02.2008

Author
AB
AB
AB
AB

Comments
Typos and minor errors fixed.
Typos and minor errors fixed.
Document put on blackboard site.
Document created.

Contents
1

The composition of stars

Gravity and hydrostatic equilibrium


2.1 Gravitational contraction . . . . . . . . . . . .
2.1.1 Free fall . . . . . . . . . . . . . . . . .
2.2 Hydrostatic equilibrium . . . . . . . . . . . . .
2.2.1 Equilibrium for a non-relativistic gas .
2.2.2 Equilibrium for an ultra-relativistic gas

.
.
.
.
.

3
3
4
5
6
7

Star formation
3.1 Conditions for gravitation collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Contraction of a protostar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Conditions for being a star . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7
7
8
9

The sun
4.1 Solar radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Nuclear fusion in the sun . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

10
10
11

Nucleosynthesis

12

The main sequence

12

Problems

14

Hints to exercises

15
1

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

REFERENCES

References
Ostlie D.A., Carroll B.W., 2007, An Introduction to Modern Stellar Astrophysics, Second Edition, AddisonWesley, San Francisco
Phillips A.C., 1998, The Physics of Stars, second edition, The Manchester Physics Series, John Wiley & Sons
ltd, Chichester

2.1

Gravitational contraction

r
r
P A

gM

(P + P )A

Figure 1: Spherically symmetric system of mass M and radius R. The arrows on the right hand side indicate
the forces acting on a small element, with volume rA located at distance r, due to gravity and pressure.

The composition of stars

Right after the big bang the process of nucleosynthesis led to the creation of the first elements in roughly the
following ratio (by mass): 75% hydrogen, 25% helium and a tiny amount of lithium-7. These elements are
the raw materials from which the first stars formed. The ratio of elements present in a star is referred to as
its chemical composition. All stars convert hydrogen to helium in their interiors and helium is subsequently
converted into the heavier elements such carbon, nitrogen, oxygen, iron, etc; the elements from which the earth
and ourselves have been built up. The elements of atomic number higher than that of helium are referred to as
metals in astronomical jargon. Stars that have formed after the first generation of stars still consist of roughly
75% hydrogen and 25% helium but also contain a small admixture of metals produced by the first stars. The
second and later generation stars are said to have a non-zero metallicity.

Gravity and hydrostatic equilibrium

The most important driving force in the formation and evolution of stars is that of gravity which leads to:
1. contraction of interstellar gas-clouds and the subsequent formation of stars;
2. conditions in the stellar interior which will allow nuclear fusion to proceed;
3. the fusion of hydrogen to helium is usually followed by further contraction of the star and the conversion
of helium (and heavier elements) to elements such as C, O, Fe, etc.

2.1

Gravitational contraction

We start by studying a number of basic properties of the compression of matter due to gravity. Consider a
spherically symmetric system of mass M and radius R as illustrated in figure 1. The only forces acting on the
matter in this sphere are its self-gravity and the internal pressure. The density and pressure at a distance r from
the centre are (r) and P (r).
The acceleration of a mass element located at a distance r from the centre of the spherical system can be
derived as follows. The mass inside the spherical shell with radius r is:
Z r
m(r) =
(r0 )4r02 dr0 ,
(1)
0

which acts as a gravitational mass located at the centre and gives rise to an inward acceleration:
g(r) =

Gm(r)
.
r2

(2)

2.1

Gravitational contraction

In general there will also be a force due to the pressure gradient. Consider a small volume element located
between radii r and r + r, with surface area A and volume rA, as in figure 1. There will be a net force
acting on the volume element if the pressure on the inside surface is different from that on the outside surface.
The inward force due to the pressure gradient is:


dP
dP
P (r) +
r P (r) A =
rA .
(3)
dr
dr
With M = (r)rA we can derive the acceleration of the volume element as:


d2 r
1 dP
.
= g(r) +
dt2
(r) dr

(4)

The acceleration is inward and thus to compensate the force of gravity the pressure gradient has to be negative.
That is, the pressure has to increase towards the centre in order to counteract the self-gravity of the spherical
system.
2.1.1

Free fall

If we now assume that there is no internal pressure we can derive how long it takes before all the mass in
the system is concentrated at the centre. From symmetry considerations it follows that every spherical shell
of matter will converge on the centre with acceleration Gm0 /r2 , where m0 is matter enclosed by the shell.
The shell is assumed to be at rest at a radius r0 and the enclosed mass is assumed to remain constant during
the collapse. From the conservation of energy (gravitation potential energy is converted to kinetic energy) it
follows that the velocity of the shell after the collapse from r0 to r is given by:
 
1 dr 2 Gm0 Gm0
=

.
(5)
2 dt
r
r0
The time is takes for a free fall towards the centre of the sphere (the free-fall time is then given by:

Z tend
Z 0
Z 0
dt
2Gm0 2Gm0 1/2
tff =
dt =
dr =

dr .
r
r0
0
r0 dr
r0

(6)

Where we take into account that dr/dt < 0 (velocity vector is directed inward). The expression between
brackets in the integral can also be written as:




2Gm0 2Gm0
2Gm0 1 r/r0
r0 r

= 2Gm0
=
.
(7)
r
r0
r0 r
r0
r/r0
With the substitution x = r/r0 the free-fall time can be written as:
 3 1/2 Z 1 
1/2
x
r0
dx .
tff =
2Gm0
1x
0

(8)

The last integral can be solved through the substitution x = sin2 and evaluates to /2. Hence the free-fall
time depends on m0 /r03 and thus on the average density within the shell with radius r0 . For a spherical system
with uniform density we can then write:


3 1/2
tff =
.
(9)
32G
In practice gravity will always be opposed to some extent because the energy released during free fall will
be converted to random thermal motions thereby building up a pressure. The free fall approximation is however
relevant if the released energy can be radiated away or used up in the excitation or dissociation of the gas
particles in the system. An example relevant to the formation of stars is the collapse of a cloud of molecular
hydrogen. It can collapse in free fall as long as it is transparent to its own radiation, or if the hydrogen molecules
are being dissociated, or while the hydrogen atoms are being ionized. As soon as the gravitational energy is
being released in a non-transparent cloud of ionized hydrogen the internal pressure will rise and the cloud will
slowly reach hydrostatic equilibrium.

2.2

2.2

Hydrostatic equilibrium

Hydrostatic equilibrium

From figure 1 and equation (4) it follows that a volume element at distance r from the centre of the spherical
system is in hydrostatic equilibrium if the pressure gradient at radius r is given by:
dP
Gm(r)(r)
=
.
dr
r2

(10)

If this equation holds at all radii then our entire spherical system is in hydrostatic equilibrium and in that case
we can find an expression which relates the average internal pressure to the gravitational energy. We multiply
both sides of the equation for hydrostatic equilibrium by 4r3 and we integrate from r = 0 to r = R:
Z R
Z R
Gm(r)(r)4r2
dP
4r3
dr =
dr .
(11)
dr
r
0
0
Both sides of the resulting equation have a physical meaning. De right-hand side is simply the gravitational
potential energy of the system:
Z m=M
Gm(r)
Egr =
dm ,
(12)
r
m=0
where dm is the mass between r and r + dr, i.e. (r)4r2 dr. The left-hand side can be evaluated through
partial integration:
Z R


3 R
P (r)4r2 dr .
(13)
P (r)4r 0 3
0

The first term in this expression is zero if we assume that the pressure outside of radius r = R is equal
to zero. De second term is equal to 3 hP i V , where V is the volume of the system and hP i the volumeaveraged internal pressure. Thus the average pressure needed to support a spherical system with volume V and
gravitational energy Egr is given by:
1 Egr
hP i =
.
(14)
3 V
In other words the average internal pressure is one-third of the density of the stored gravitational energy. This
expression for the average pressure needed to support a self-gravitating system is one form of the so-called
virial theorem.
To examine this result further we will now consider the pressure generated by a classical gas due to the
translational motion of the gas particles. Consider a gas of N particles confined to a cubical box with volume
L3 with its edges oriented along the x, y and z axes. Take one gas particle with velocity v = (vx , vy , vz )
and momentum p = (px , py , pz ). As this particle moves about in the box it bounces off the walls at regular
intervals. The rate at which it strikes one of the sides perpendicular to the z-axis is vz /2L. Upon bouncing off
the wall the particle will impart a momentum 2pz (as follows from momentum conservation with the wall being
at rest). So the rate of momentum transfer to a unit area of the wall is pz vz /L. If now consider all particles in
the box then the pressure generated on a wall of area L2 perpendicular to the z-axis is:
P =

N
hpz vz i ,
L3

(15)

where the brackets denote an average over all particles. If we assume the gas to be isotropic then all directions
of motion for the particles are equally likely and:
hpx vx i = hpy vy i = hpz vz i =

1
hp vi ,
3

(16)

where
p v = px vx + py vy + pz vz .

(17)

Hence the pressure of each side of the box is the same and given by:
P =

n
hp vi ,
3

(18)

2.2

Hydrostatic equilibrium

where n is the number of particles per unit volume. Note that this expression is valid for classical gases and
for gases in which quantum effects are important. In addition it is valid for both non-relativistic and relativistic
gases.
We now consider the pressure for two types of ideal gases, one with non-relativistic particles and one with
ultra-relativistic particles. The general expression for the energy p for a particle of mass m with momentum p
is:
2p = p2 c2 + m2 c4 ,
(19)
where the velocity of the particle is v = pc2 /p . De non-relativistic limit follows from the assumption p  mc
for which p = mc2 + p2 /2m and v = p/m. The ultra-relativistic limit follows from p  mc so that p = pc
and v = c. The general expression for the pressure in an ideal gas derived above has the following form for
these two limits:
for a non-relativistic gas p v = mv 2 and the pressure is:


2
1
2
P = n
mv
,
3
2

(20)

or 2/3 of the translational kinetic energy density.


for an ultra-relativistic gas p v = pc and the pressure is:
1
P = n hpci ,
3

(21)

or 1/3 of the translational kinetic energy density.


2.2.1

Equilibrium for a non-relativistic gas

For a non-relativistic self-gravitating ideal gas with volume V we can write for the average pressure:
hP i =

2 Ekin
,
3 V

(22)

where Ekin is the total translational kinetic energy of the gas particles. A comparison with the virial theorem
shows that for a non-relativistic gas in hydrostatic equilibrium the following condition holds:
2Ekin + Egr = 0 .

(23)

The total energy Etot is equal to Ekin + Egr if the particles have no internal degrees of freedom (i.e. they cannot
rotate, vibrate, etc). In that case we can write:
Etot = Ekin

and

1
Etot = Egr .
2

(24)

These equations are very important in astrophysics and the describe the consequences of the virial theorem for
a system of non-relativistic gas particles that are held together by gravity and are in hydrostatic equilibrium. A
couple of remarks:
For a bound system we have Etot < 0 and the stronger the system is bound the more negative the value
of Etot is. This can be seen be rewriting equation (23) as Egr = 2Ekin , or |Egr | > Ekin , which means
that Ekin + Egr < 0. The quantity Etot is the binding energy of the system.
The binding energy of the system, Etot , is equal to the total kinetic energy. Hence strongly bound
clouds of gas are hot!
If the system slowly changes but stays near hydrostatic equilibrium then the changes in the kinetic and
gravitational energies are related in a simple way to the change in total energy. A decrease of Etot by 1%
leads to a decrease of Egr by 2% and in increase of Ekin by 1% (so that equation 23 is still satisfied).

3.1

Conditions for gravitation collapse

This kind of behaviour often occurs in astrophysical systems. Consider a gas cloud which loses energy from
its surface by radiation. If this energy loss is supplied by the release of gravitational energy then the gravitational
energy decreases while the kinetic energy increases, that is the cloud contracts and heats up. If this contraction
occurs near hydrostatic equilibrium half the released gravitational energy is lost through radiation and the other
half is used to heat up the gas. This heating up in turn causes an increase in pressure which is to compensate
the increased gravity. However, the continuing energy loss will cause the cloud to keep contracting.
The cloud contraction can be halted if the energy loss at its surface is compensated by, for example, energy
released in nuclear fusion processes in its interior. This is what the sun does. If more energy is released
through nuclear fusion that is radiated away the cloud will expand and cool (the total energy increases thus
Ekin decreases and Egr increases). Conversely, if the nuclear reaction cost energy the energy loss of the system
will increase (energy is lost by radiation at the surface and absorbed by the nuclear reactions) and therefore the
system will contract and the gas will heat up.
Energy loss for stars thus leads to a heating up of the interior and an energy increase leads to cooling of the
interior. This means that the specific heat capacity of stars is negative!
2.2.2

Equilibrium for an ultra-relativistic gas

For an ideal gas with ultra-relativistic particles the pressure is 1/3 of the kinetic energy density and we can
write:
Ekin + Egr = 0 .
(25)
This means that hydrostatic equilibrium is only possible if the total energy is zero. Such a system is between
bound and unbound and can easily be disrupted. This kind of instability occurs in stars for which the pressure
is generated primarily by photons (an ultra-relativistic gas). This instability can also occur in stars with a
degenerate electron-gas in the interior when the electrons become very energetic.

Star formation

Stars are formed from clouds of molecular hydrogen (H2 ) and most stars from in groups, named clusters: There
are broadly speaking three types of clusters:
Globular clusters consisting of many thousands to hundreds of thousands of stars. They are generally very
old (more than a few Gyr) and the stars contain very little metals (i.e. they are metal-poor).
Open clusters consisting of 501000 stars which are relatively young (< 1 Gyr) and metal-rich.
Associations very loose groupings of stars which are very young (< 50 Myr) en often still surrounded by the
remnants of the clouds from which they formed.
In the rest of this section we will consider very global properties of star formation from elementary physical
considerations from we will see that it is plausible that stars should form in groups.

3.1

Conditions for gravitation collapse

Under which conditions can a cloud of gas collapse by its own gravity in order to form stars? The cloud
should be compact enough for its mass such that gravity will overcome the internal pressure, or |Egr | > Ekin .
Consider a cloud of mass M and radius R consisting of N particles of average mass m. The cloud has a uniform
temperature T . The gravitational energy can be calculated using equation (12) and equals:
Egr = f

GM 2
,
R

(26)

where the factor f depends on the density distribution in the cloud. For a spherical cloud with uniform density
it can be shown that f = 3/5. For the rough estimates that follow we will use f = 1. The kinetic energy can
be written as:
3
Ekin = N kT .
(27)
2

3.2

Contraction of a protostar

The condition for the collapse of the cloud is |Egr | > Ekin from which it follows that a cloud of radius R has
to have a mass larger than:
3kT
MJ =
R.
(28)
2Gm
We can also write this as a density criterion for a cloud of mass M . The condition for the radius is:
R<

GM 2
2Gm
=
M,
3
3kT
2 N kT

(29)

from which it follows (with = 3M/(4R3 )) that the density has to be larger than:
3
J =
4M 2

3kT
2Gm

3
.

(30)

The subscripts J in these expressions indicate that the criteria define the so-called Jeans mass and density.
The expression for the Jeans density shows that it has to be low in order to make it easy for a cloud to
collapse. This is easier for a cloud with a large mass. For example, a cloud with a mass of 2 1033 kg (1000
M ) can collapse if the density becomes larger than 1022 kg m3 , about 105 molecules per cubic meter. For
a cloud of 1 M the Jeans density is one million times higher!
From these considerations it follows that the process of star formation may proceed as follows. First a large
and massive cloud will start to contract by its own gravity. As the density becomes high enough smaller pieces
of the cloud in the interior can start to collapse independently. The cloud will thus fragment into smaller clouds
from which eventually protostars can form. This gives a qualitative explanation as to why we can expect stars
to form in groups.

3.2

Contraction of a protostar

The expression for the Jeans density indicates that if a cloud at a temperature of 20 K can reach densities of
about 1016 kg m3 it can fragment into smaller clouds of 1 M which can contract independently. Such a
fragment will constitute a protostar with a radius of about 1015 m, 106 R . As long as the gravitational energy
being released is not converted into thermal motion of the molecules (pressure) the fragment will collapse in
free fall. This can go on as long as the dissociation of H2 and the ionization of H absorb the released potential
energy. The dissociation and ionization energies are D = 4.5 eV and I = 13.6 eV. So the energy required to
convert all the H2 to HII is roughly:
M
M
D +
I ,
(31)
2mH
mH
where mH is the mass of the hydrogen atom. This energy is released during the contraction of the protostar
from a radius R1 to a radius R2 , hence:
M
M
GM 2 GM 2

D +
I .
R2
R1
2mH
mH

(32)

In this example (protostar of 1 M ) the required energy is 3 1039 J and the protostar will contract in free
fall from R1 1015 m to R2 1011 m, or about 100 R . The timescale is given by the expression for tff
(equation 9) with 1016 kg m3 and is about 200 000 yr.
After the complete ionization of all hydrogen (in the interior) the contraction of the protostar will proceed
until the gas becomes opaque to its own radiation and then the released gravitational energy will be converted
into random thermal energy of the electrons and protons. The pressure will rise in this process and the protostar
will slowly reach hydrostatic equilibrium. With the aid of the virial theorem we can estimate the internal
temperature at the moment the contraction starts to slow down. De kinetic energy of the electron and protons
in the gas is:
M
Ekin
3kT ,
(33)
mH

3.3

Conditions for being a star

where we have used m 0.5mH . Because R1  R2 we can write for the gravitational energy:


M
M
GM 2

D +
I .
Egr
R2
2mH
mH

(34)

According to the virial theorem we have 2Ekin + Egr = 0 thus for the temperature we can write:
kT

1
(D + 2I ) 2.6 eV .
12

(35)

This amounts to a temperature of approximately 30 000 K. This estimate is independent of the mass of the
protostar.
The protostar will now continue to contract slowly thereby maintaining approximate hydrostatic equilibrium. According to the virial theorem half the released potential energy is converted to internal kinetic energy
and the rest is lost as radiation at the surface. The internal pressure and temperature now will continue to
increase until the conditions are suitable for the thermonuclear fusion of hydrogen to helium. The nuclear reactions will release energy which will compensate the energy loss at the surface and then the contraction of the
protostar stops and it becomes a real star.

3.3

Conditions for being a star

Not all contracting protostars will in the end become stars. De pressure needed to stop the contraction can also
be generated by a cold, dense gas of degenerate electrons. In such a gas the behaviour of electrons is governed
by the laws of quantum mechanics and the electrons will occupy the lowest possible energy states in accordance
with the Pauli exclusion principle. Such a gas generates pressure not by the increasing random thermal energy
of the electrons but because the total kinetic energy of the electrons has a minimum which increases as the
density rises. In fact the temperature no longer rises.
An electron gas becomes degenerate if the average distance between the electrons is comparable to the
the Broglie wavelength = h/p. The momentum of electrons in a classical electron gas at temperature T is
roughly (me kT )1/2 (kinetic energy of kT ) and then the typical the Broglie wavelength is:
=

h
.
(me kT )1/2

(36)

The electron gas is a classical gas as long as the wave functions of the electrons do not overlap, i.e. the average
mutual distance between the electrons must be much larger than . The density of the gas then has to satisfy
the following condition:
m
(me kT )3/2
,
(37)
 3 m

h3
where m = 0.5mH for a gas with an equal amount of electrons and protons. As long as the gas is classical the
temperature will rise as the density in the protostar increases. Using the approximate expression for Egr , the
classical expression for Ekin , and the condition for hydrostatic equilibrium (equations 26, 27 and 23) we find:
kT

GM 2
GM m
=
GmM 2/3 1/3 ,
3N R
3R

(38)

where we have used M/3R M 2/3 1/3 . Thus while the density remains below the critical value the temperature will rise with 1/3 . The temperature at which the electrons become degenerate can be estimated by
substituting the critical density in the expression for kT above:
kT GmM 2/3 m1/3

(me kT )1/2
,
h

or
kT

G2 m8/3 me
h2

(39)

!
M 4/3 .

(40)

4.1

Solar radiation

10

Hence the mass of the protostar determines the internal temperature that is reached before the electron gas
becomes degenerate! For 1 M the temperature reached is kT 1 keV. In other words if a protostar with
a mass of 1 M keep contracting by its self-gravity eventually an average internal temperature of about 10
million K is reached, with a central temperature which is even higher. This is more than sufficient to trigger
nuclear reactions and the conversion of hydrogen to helium. For less massive stars lower temperatures are
reached (making it more difficult to ignite nuclear reactions). Detailed calculations show that in objects less
massive than 0.08 M nuclear reactions will not be triggered. These are protostars that evolve toward a state in
which gravity is opposed by the pressure of a degenerate electron gas, the so called brown dwarfs.
At the other end of the mass spectrum we find the very massive stars for which the internal pressure is
dominated by photons and we have already seen that for such an ultra-relativistic gas the equilibrium is unstable.
Above 50100 M stars are expected to easily disrupt and these are indeed very rare.

The sun

Because our sun is the best known star for which we have the most accurate information we will proceed to
apply to it a number of the estimates we have derived about the interiors of stars.
The sun has a mass of about 2 1030 kg and a radius of about 7 108 m. Hence the average density hi is
1.4 103 kg m3 , comparable to the density of water. The free-fall time then is only a half hour according to
equation (9). We can thus safely conclude that the sun is not in a state of free fall and that the internal pressure
must be enough to oppose gravity. The age of the sun is about 4.5 billion years and we have no evidence that
during its lifetime any drastic changes occurred (e.g., from the geological record). Hence the sun has remained
in hydrostatic equilibrium during all that time which means we can determine the average internal pressure
using the virial theorem:
GM 2
1 Egr
14
hP i =

(41)
4 10 Pa ,
3 V
4R
which is a billion times the pressure in the earths atmosphere. Despite this high pressure we can still treat the
solar interior as a classical ideal gas and that case the average internal pressure is also given by:
hP i =

hi
kTin ,
m

(42)

where Tin is the typical internal temperature and m = 0.5 amu (atomic mass unit) for a gas of ionized hydrogen.
The standard model of the sun in fact has a composition of 71% H, 27% He, and 2% heavier elements for which
m = 0.61 amu. We can now estimate the internal temperature as:
kTin

GM m
0.5 keV
3R

or

Tin 4.6 106 K .

(43)

The real central density, temperature and pressure are: 1.48 105 kg m3 , 15.6 106 K, 2.29 1016 Pa.
So our estimates are pretty good!

4.1

Solar radiation

The sun radiates about 4 1026 W of energy an has a corresponding effective temperature of about 6000 K.
Because kTeff 0.5 eV most of the radiation comes out in the visual. Hence the temperature at the surface of
the sun is about 1000 times lower than in the interior. If the radiation could leave the solar interior unhindered
its luminosity would be:
2
4
L0 4R
Tin
,
(44)
in which case the sun would radiate primarily in the X-ray regime (kTin 0.5 keV). Luckily this is not the
case but what happens to the energy produced in the interior?
The radiation from the interior of the sun is constantly scattered, absorbed and emitted by the electrons and
ions in the gas. A temperature gradient is set up and the radiation slowly diffuses to the surface. The photons

4.2

Nuclear fusion in the sun

11

diffuse according to the so-called random walk. If l is the average free path length for a photon then the
displacement after N interactions is with the gas is:
D = l1 + l2 + + lN ,

(45)

where we assume that l is constant over the whole solar interior. The value of D2 is:
2
D2 = l12 + l22 + lN
+ 2(l1 l2 + l1 l3 + )

(46)

For a random walk the second term averages to zero so the average of the square of the displacement after N
interactions is N l2 . To escape the sun an average photon has to bridge a distance comparable to the solar radius
2 /l2 steps. The time for each step is about l/c so the random walk escape time is:
and this takes about R
trw

2
R
cl

(47)

The time it takes to escape the sun directly is R /c which is a factor l/R lower than trw . Hence the rate at
which energy escapes the sun is lower by the same factor as a consequence of the diffusion process. This means
that the luminosity is also lower by a factor l/R compared to when the radiation can leave the sun freely. So
for the effective temperature we have:


l 1/4
Teff
Tin .
(48)
R
From this result we can conclude that the mean free path for a photon in the sun is about 2 mm. The sun is
very opaque. According to the equation for trw it takes a photon 30 000 years to diffuse from the centre to the
surface!
Using the relation between Teff and Tin and the relation between the internal temperature and the mass and
radius of the sun (eq. 43), we can write:
2
4
L 4R
Tin

l
(4)2 4 4

G m hi lM 3 .
R
35 k 4

(49)

Thus we expect that for stars like the sun the luminosity steeply increases with mass!
The diffusion process for photons restricts the flow of energy and prevents the sun from catastrophically
losing energy. This process determines the amount of energy that has to be released in nuclear reactions to
compensate the energy loss at the surface.

4.2

Nuclear fusion in the sun

The nuclear fusion process in the sun proceeds according to the so-called proton-proton chain:

p + p d + e+ + e ,

(50)

p + d 3 He + ,

(51)

He + 3 He 4 He + p + p ,

(52)

where d indicates the deuteron. Each of these reactions is exothermic and in total an energy of 26 MeV is
released for each 4 He nucleus formed.
All these reactions are in principle hindered by the Coulomb barrier and can only proceed due to quantum
mechanical tunnelling. The first reaction is the slowest and on average it takes 5 billion years for a proton to
fuse with another proton to a deuteron. The deuteron is converted to 3 He within 1 second and the time required
to convert two 3 He nuclei to 4 He is about 300 000 years. So the first reaction determines the rate at which
energy is generated within the sun. The luminosity of the sun is enormous but on average every kilo in the sun
only produces 0.2 mW, which is about 10 000 times less than the heat we produce per kilo in our bodies!

The main sequence

12

Table 1: The most important stages of thermonuclear fusion in stars. The ashes of one stage of burning form
the fuel for the next stage as long as the contracting star is massive enough to reach the temperatures indicated.
Process
Fuel
Products
Ignition temperature
(K, approximate)
Hydrogen burning Hydrogen Helium
1 107
Helium burning
Helium
Carbon, oxygen
1 108
Carbon burning
Carbon
Oxygen, neon, sodium, magnesium
5 108
Neon burning
Neon
Oxygen, magnesium
1 109
Oxygen burning
Oxygen
Magnesium to sulphur
2 109
Silicon burning
Silicon
Iron and nearby elements
3 109

To produce one 4 He 4 protons are used and 26 MeV of energy is released. Hence the rate at which protons
are used up is:
(4 4 1026 )/(26 1.6 1013 ) = 4 1038 protons per second .
(53)
In this process at least 2 1038 neutrinos are released which can leave the sun unhindered.
The nuclear reactions in the suns interior act as a kind of thermostat. If the temperature in the interior
rises more nuclear reactions will occur and the total energy will increase (the nuclear reactions producing more
energy than can escape the sun). From the virial theorem we know that the sun will then expand and cool.
Conversely, a lowering of the internal temperature will lead to a contraction and heating up of the sun. This
thermostat has functioned for the past 4.5 billion years and will do so until there is no longer enough hydrogen
at the centre of the sun to keep the proton-proton chain going. There are about 7 1056 protons in the sun and
about 10% of those will be converted to 4 He in the next 6 billion years. The total hydrogen burning phase for
the sun lasts about 10 billion years. After that time the core will contract and heat up until the pressure and
temperature are high enough for the burning of helium to start. At that moment the outer layers of the sun will
expand and cool and the sun becomes a red giant.

Nucleosynthesis

Stellar evolution is a process whereby gravitational energy is released by the contraction of the star. This
contraction temporarily stops if enough energy is added to the system by nuclear reactions so that the energy
loss at the surface is compensated. The ashes of one nuclear burning stage form the fuel for the next stage.
There is in fact a sequence of thermonuclear burning stages at ever increasing temperatures. This is described
in table 1.
The process of nuclear fusion can only continue as long as the formation of a new nucleus produces net
energy. In figure 10.9 from Ostlie & Carroll (2007) the binding energy per nucleon is plotted. Starting from
hydrogen we see that as long as the binding energy increases nuclear fusion will produce net energy. This stops
as soon as iron is reached. The elements near iron are formed (Cr, Mn, Fe, Co, Ni) but constitute an ash which
cannot be burned any further and thus they constitute the end stages of nuclear fusion processes in the interiors
of stars.
Keep in mind that the particular nuclear burning stage that can be reached by a star depends on the central
temperatures reached which in turn depend on the mass of the star. Only the most massive stars will eventually
produce iron in their cores. Elements heavier than iron can only be formed through the capture of neutrons and
-decay. This only occurs during the very last stages of stellar evolution.

The main sequence

Stars that convert hydrogen to helium in their interiors are in the phase of their evolution in which they are
located on the main sequence in the Hertzsprung-Russell diagram. We have seen that the luminosity of a star
is a steeply increasing function of its mass (eq. 49). If we assume that the mean free path l is inversely

106

The main sequence

13

Main sequence massluminosity relation

Luminosity (LSun)

104

102

100

102
0.1

1.0

10.0

100.0

Mass (MSun)
Figure 2: The main sequence mass-luminosity relation according to the data listed in appendix G of Ostlie &
Carroll (2007). The line is the estimated relation L M 3 .
proportional to the average density we expect that for stars on the main sequence L M 3 . Figure 2 shows the
actual relation between mass and luminosity for stars on the main sequence according to the data in appendix
G of Ostlie & Carroll (2007). The real relation is close to our estimate. In reality the relation L M holds,
with 3 for massive stars and 3.5 for stars less massive than the sun.
De rapid increase of luminosity with mass means that more massive stars burn their nuclear fuel faster
than less massive stars and thus have shorter main sequence lifetimes. For massive stars the main sequence
lifetime is proportional to M 2 while for less massive stars the lifetime is proportional to M 2.5 . The sun has
a main sequence lifetime of 10 billion years so a star of 10 M only spends about 10 million years on the main
sequence while a star of 0.5 M is on the main sequence for more than 50 billion years!
The mass of a star is by far the most important parameter determining its evolution.

Problems

14

Problems

The problems below are taken from the book by Phillips (1998). Try to solve the problems first without consulting the hints!
1. Consider a sphere of mass M and radius R.
a. Calculate the gravitational potential energy of the sphere assuming (A) a density which is independent of the distance from the centre, and (B) a density which increases towards the centre according
to:
(r) = c (1 r/R) .
b. In both cases, (A) and (B), write down the average internal pressure needed for hydrostatic equilibrium, and
c. determine how the pressure in the sphere depends on the distance from the centre.
2. As the sun evolved towards the main sequence, it contracted under gravity while remaining close to
hydrostatic equilibrium, and its internal temperature changed from about 30 000 K, eq. (35) to about
4.6 106 K, eq. (43). (This stage of stellar evolution is called the Kelvin-Helmholtz stage.)
a. Find the total energy radiated during this contraction. Assume that the luminosity during this contraction is comparable to the present luminosity of the sun and estimate the time taken to reach the
main sequence.
b. Explain how it is possible that while loosing energy the contracting sun heats up.
3. The main sequence of the Pleiades cluster of stars consists of stars with mass less than 6 M ; the more
massive stars have already evolved off the main sequence.
a. Estimate the age of the Pleiades cluster.
b. The real age is 120 Myr. Explain why the estimated age is different.
4. Useful bounds can be set on the pressure at the centre of a star without detailed stellar structure calculations. Consider a star of mass M and radius R. Let P (r) be the pressure at distance r from the centre
and m(r) be the mass enclosed by a sphere of radius r. Show that in hydrostatic equilibrium the function
P (r) +

Gm(r)2
8r4

decreases with r. Hence show that the central pressure satisfies the inequality
1
Pc >
6

4
3

1/3

G hi4/3 M 2/3 ,

where hi is the average density.


If you assume that the density (r) decreases with r, it is possible to derive a tighter lower bound and, in
addition, a useful upper bound for the central density. Show that
Pc >

1
2

4
3

1/3

In addition show that


1
Pc <
2
where c is the central density.

4
3

G hi4/3 M 2/3 .

1/3

2/3
G4/3
,
c M

Hints to exercises

15

Hints to exercises
1. Use equations (12), (14), and (10).
3. Bear in mind that the luminosity of a star of mass M is proportional to M with between 3 and 3.5.
4. Let

Gm(r)2
df
and show that
< 0.
4
8r
dr
The first lower bound on Pc is given by the condition f (0) > f (R). The second lower bound and the
upper bound on Pc can be obtained by noting that
f (r) = P (r) +

m(r) >

4
hi r3 ,
3

m(r) <

and using
df
Gm(r)2
= 0.
+
dr
2r5

4
c r3 ,
3

Das könnte Ihnen auch gefallen