Download Section C - Astrophysics

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

IK Pegasi wikipedia , lookup

R136a1 wikipedia , lookup

Future of an expanding universe wikipedia , lookup

Star formation wikipedia , lookup

P-nuclei wikipedia , lookup

Stellar evolution wikipedia , lookup

Degenerate matter wikipedia , lookup

Transcript
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
(Section C) The Internal Structure of Stars
Stars are large spheres of hot gas in equilibrium between internal pressure and selfgravity. They are then governed by the law of hydrostatic equilibrium (see
Astrophysical Concepts)
Hydrostatic Equilibrium
[Figs. 38-40: Hydrostatic Equilibrium]
Assume that the star is spherically symmetric so that pressure P, density  and
temperature T are all functions of radius r only.
If we consider a fluid element of cross-sectional area A and depth r at radial
position r, then the outward force due to the pressure (gradient) will be just
Fp = PA - (P + P)A = - PA = - r(dP/dr)A.
(Note that the force is outward only if dP/dr is negative).
The inward gravitational force on the element (mass m), due to the mass inside
GM ( r )m
radius r will be
F 
g
r2
But given the respective volumes and densities, the masses must be m = (r)rA
and M ( r )  r  ( s )4s 2 ds
0
For equilibrium Fp = Fg so dP   GM ( r )  ( r )
2
dr
r
To solve this explicitly, we need to know (or assume) (r) or have another
relationship between P and . However, we can obtain some simple approximations.
Purely on dimensional grounds it is clear that if we just approximate dP/dr by the total
change in pressure from the centre of the star (P = Pc) to the surface (P = 0) divided by
2
the radius of the star (R), of total mass M, we must have P  GM  3GM
c
R
4R 4
More accurately, assume a constant density   3M 4R 3 . This implies
M(r)=(r/R)3M, so putting these into the hydrostatic equilibrium equation we have

dP GMr 3 3M 1



dr
R 3 4R 3 r 2
Integrating from the centre to the surface we then get
R
Pc  0 rdr
3GM 2
4R 6

3GM 2
8R 4
For the Sun, M = 2 x 1030kg, R=7 x 108m (so  = 1400 kg m−3), giving
Pc ≈1.4 x 1014 Nm−2. For a realistic (r) decreasing outwards, Pc is even higher. Such
pressures must therefore exist at the centres of stars to support them against their own
self gravity. We should next consider how these pressures are produced.
Page 1 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
Thermal Gas Pressure
In a gas at temperature T the particles (mass m) have mean kinetic energy
(1/2)m <v2>= (3/2)kT.
Momentum transfer through collisions then produces a pressure given, for an ideal
gas, by the perfect gas law
P=nkT=(k/m)T
where n and  are the particle number density and the mass density, respectively.
The relationship between P and ρ (and in this case T) is called the equation of state.
Provided we have an ionized gas (plasma), the perfect gas law still holds at the very
high densities calculated above.
Assuming a star to be made entirely of ionized hydrogen the mean particle mass
m = (me + mp)/2 ≈ mp/2. (If we include helium this increases to about 0.62 mp). Using
the central pressure calculated for a homogeneous sphere and the perfect gas law we
get a central temperature
Tc 
m Pc m GM
  
 6 10 6 K
k  k 2R
Note that at these temperatures
(3/2)kT ~1 keV >> 13.6 eV, so the hydrogen is indeed ionized. They are also high
enough for thermonuclear fusion reactions to occur.
Radiation Pressure
A photon with energy hν carries momentum hν/c and therefore exerts pressure
through the transfer of this momentum when photons are absorbed or scattered. For a
black body we know that the flux (of energy of photons) passing through unit area of
a surface is σT4. Thus we would expect a pressure (flux of momentum) ~ σT4/c.
Specifically, for an isotropic radiation field this becomes
Prad
4 T 4 1 4
 
 aT
3 c
3
At T~107 K. Prad ~3×1012 Nm-2 ~ Pc/50
However, as T increases, Prad may dominate.
Page 2 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
Degeneracy Pressure
Degeneracy pressure is a quantum mechanical effect, important at very high densities.
It occurs because the Pauli exclusion principle allows only one electron per quantum
state (or 2 per energy state, as the electron has a choice of two spin states). The
number of available states per unit volume is limited for a confined set of particles.
The exclusion principle thus prevents the particles being too close together and
therefore acts as a pressure.
The Heisenberg uncertainty principle relates position (x) and momentum (p)
uncertainties (in 1-D) through x p> h. If a particle is constrained in a very small
volume then its momentum must be large.
In 3-D this means that we can only have 2 electrons in a 6-dimensional phase-space
element defined by 4r2dr × 4p2dp = h3.
Thus the (physical) space density of particles with momentum p to p + dp is limited
to dn = 2/4r2dr = 8p2dp/h3.
The total density of particles with momentum less than p is then just the integral of
this, i.e. n=(8p3)/(3h3) for completely degenerate gas.
As before, pressure is the same as rate of momentum transfer, pnv  p2n for nonrelativistic particles or pnc  pn for relativistic particles.
But from above, n  p3 so we have a degeneracy pressure Pd  p5 in the nonrelativistic case or  p4 for the relativistic one.
Finally, since ρ  n  p3, this implies that non- relativistic degenerate matter will
have Pd = A1ρ5/3 while relativistic degenerate matter has Pd = A2ρ4/3 for some
constants A1 and A2.
Notice particularly that there is no dependence on temperature. Thus degenerate stars
will not change their size as they cool.
Electron degeneracy pressure is important in the very dense white dwarf stars. Since
collisions tend to equalise the energy (mv2) of particles in a gas, the much more
massive nuclei will have higher momentum (mv) by a factor (mn/me). They will
therefore occupy a (mn/me)3/2 times larger volume in momentum space than the
electrons, so do not contribute much to the pressure. At very high densities, proton
and electrons combine to form neutrons (inverse beta-decay). Neutron degeneracy can
set in when the number density is about 105 times higher than for electron degeneracy.
This occurs in neutron stars.
[Fig. 41 : Condensed Objects]
Page 3 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
(Section D) Energy Sources in Stars
As a star shines it is losing energy at a rate equal to its luminosity L. If it has a
lifetime τ it must have available an energy source E=τL.
From estimates of the age of the Earth, we know that the Sun must be at least 4.6 Gyr
old (1.5 × 1017s). This requires E > 6 × 1043J.
Consider first gravitational energy. As the separation of two masses m1 and m2 is
reduced from infinity to r, gravitational potential energy is released :-
EG 
GM 1M 2
r
In the same way, if a star of mass M is formed by the collapse of a
very large cloud down to radius R the energy released is E   GM
G
r
constant which depends on the density distribution of the star.
2
where  is a
If the density is uniform,  = 3/5. Thus for the Sun we have an approximate energy
supply :-
3 (6.67  10 11 )( 2  10 30 ) 2
EG  
 2  10 41 J
8
5
7  10
This is about 300 times too small.
Or, put another way, the lifetime due only to gravitational contraction (the KelvinHelmholtz timescale) is 300 times too short.
Nuclear Energy
A general atomic nucleus consists of Z protons and N neutrons. Z is the atomic
number and the number of nucleons, A = Z + N, is the mass number.
For light nuclei, N ≈ Z.
In the nuclear shell model, the protons and neutrons fill up parallel sets of energy
levels, in a similar way to the electrons in shells outside the nucleus. The binding
energy of a nucleus is
A
B  ( Zm  Nm  M A )c 2 where M Z is the total mass of the nucleus. B is usually
p
n
Z
given in MeV and is the amount of energy needed to pull the nucleus apart. For
comparing the stability of nuclei, the binding energy per nucleon B/A is more useful.
[Fig. 42: Binding Energy]
Two forces are important in nuclei, the strong nuclear force and the electrostatic
force. The strong force is attractive and applies between all nucleons. As its name
implies it is stronger than other forces on nuclear scales, but is very short
range,~10−15m (smaller than one large nucleus). The strong force also saturates, in the
sense that a nucleon can only bind to a limited number of others around it. In dense
nuclear matter all bonds will be saturated so we would expect B/A to tend to a
constant value, as seen for heavy nuclei. On the other hand, for light nuclei, there are
Page 4 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
unused bonds so additional nucleons can be pulled in closer. B/A should therefore
increase with A.
The electrostatic (Coulomb) force is repulsive (all positive charges) but affects only
the protons. It is weak but extends to all scales. Its contribution to the binding energy
works like gravitational potential,
3 Z 2 e 2 for a nucleus of size R.
B 
5 4 0 R
The density of nucleons is roughly constant so R  A1/3. Also, until we reach very
heavy nuclides Z ≈A/2. Therefore
B
Z2
  4 / 3   A2 / 3
A
A
In heavy nuclei Z does not increase as fast as A, so B/A is reduced. The opposing
effects of the two forces produce a peak in B/A for moderately heavy nuclei,
specifically at A = 56, i.e. Iron. Fe56 is the most stable nucleus.
This implies that energy will be released if nuclei up to Fe are synthesised from
lighter ones (nuclear fusion) and will also be released if nuclei heavier than Fe are
split into two lighter ones (nuclear fission).
[Fig. 43: Binding Energy]
Stars are mostly hydrogen and helium, so fusion is potentially important for supplying
their energy output but fission is not. For fusion to occur, the electrostatic repulsion
between the two light nuclei must be overcome.
This is possible in stars because of the high thermal energies of the nuclei. This leads
to thermonuclear fusion. (Actually, quantum tunnelling through the coulomb barrier is
crucial to this process which would otherwise require temperatures some 1000 times
higher.) Bigger nuclei have more coulomb repulsion between them, so higher
temperatures are required for fusion. This is important for understanding stellar
evolution. The most important fusion reactions in stars are as follows.
Hydrogen Burning
The net effect of hydrogen burning is to turn four protons into a helium nucleus
4H1 1→ He42+ 2e+ + 2νe
where e+ represents a positron and νe is an electron neutrino. This occurs at a
temperature TH ~ 107K and is the energy source for main sequence stars; the net result
in each case is the release of 26.7 MeV per He4 nucleus produced.
[Fig. 44: Fusion: 4p->He]
However, it does not proceed by the direct fusion of four protons (as it is unlikely that
four will all arrive at the same point at the same time), but by successive addition.
Several chains of reactions are important.
The first are the PP (proton-proton) chains. In this case He4 is built up via deuterium
and He3 (PP I), or via the production of Be7 and then either Li7 (PP II) or B8 (PP III).
Page 5 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
[Fig. 45: Fusion details] [Fig. 46: Fusion reactions] Fig. 47: More PP reactions –
see below]
+
1
H + 1H  2H + e +  e
2
H + 1H 3He + 
3
He + 3He  4He +1H + 1H (PP I)
OR
3
He + 4He  7Be +  (PPII and PPIII)
PP Chain
3
He + 4He  7Be +
7
Be + e  7Li + 
7
Li + 1H  4He +4He
-
PPII reactions
7
Be + 1H  8B + 
8
B  8Be + e + 
8
Be  4He + 4He
+
PPIII reactions
Page 6 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
PP I dominates at T < 1.4 × 107K and only 2% of the available energy is lost to the
neutrinos (average energy 0.26 MeV). PPII dominates at 1.4 to 2.7 × 107K and has
losses of 4% to the neutrinos (the average energy of the neutrino arising from the
production of Li7 is 0.8 MeV). Finally, the PP III chain dominates above 2.3 × 107K,
but with 28% of the energy going into neutrinos (7.2 MeV for the neutrino emitted by
the Be8 decay).
The CNO cycle (sometimes called the CNO bicycle) is rather different in that it uses
existing carbon nuclei in the build up of the helium, though they are not ‘used up’,
being returned at the end after processing through oxygen and nitrogen isotopes.
Neutrinos produced at various stages all have energies ~1 MeV so the net energy loss
is 6.4%.
[Fig. 48: CNO cycle] [Fig. 49: CNO cycle reactions]
Since 26.7 MeV is produced per 4 protons, this amounts to 0.7% of the original rest
mass 4mpc2. The energy available from 1 M is therefore
Enuc = M/4mp ×26.7 × 1.6 × 10-13 ≈ 1.3 × 1045J.
Even if only 10% of the mass of a star like the Sun is used up in hydrogen burning on
the main sequence this is still enough to fuel it for ≈1010 years, easily covering the
known age of the Solar System.
Helium Burning
The next basic process is the conversion of three helium nuclei to one carbon. The
first two He42 give a Be84 and although most of these decay back
(half-life ≈ 2.6×10-16 sec), a few live long enough to fuse with another He42 to give
C126 . This is the ‘Triple-alpha’ reaction. This operates at temperatures T ~ 2 × 108K
and is a main energy source in red giants, producing 7.65 MeV per C nucleus formed.
[Fig. 50: Helium and Carbon burning]
Carbon Burning
The simplest reaction is just the fusion of two C12 6 to give Mg24 12. It requires much
higher temperatures, T~2 × 109K in order to instigate. It therefore only occurs in
massive stars, late in their lives.
There are also carbon burning processes at slightly lower temperatures (> 8 × 108K)
which produce oxygen, neon and sodium.
Note that these differ from the CNO cycle in that the carbon is used up as fuel, not
merely used as a catalyst. Even heavier elements (up to iron) can be produced in
nuclear fusion at yet higher temperatures in higher mass stars, viz.
Carbon burning (C producing O, Ne) at 5 × 108K (M> 8M)
Neon burning (Ne producing O, Mg) at 1.5 × 109K (M > 5M)
Oxygen burning (producing elements up to S) at 2 × 109K (M> 10M)
Silicon burning (producing elements up to Fe) at 3 × 109K (M > 11M)
Elements heavier than iron are produced by successive neutron capture in red giants
and supernova explosions.
Page 7 in tranche 3 of 8 pages. There were 18 previous pages
STARS Level 1 2006-07
Sections C and D
M. R. W. Masheder
Stability of Energy Generation
In equilibrium, we require that the power generated in the core is equal to that
radiated from the surface. But the energy generation rate is very temperature
dependent, so the temperature in the core must self-regulate.
In normal stars (supported by thermal pressure), if the energy generation rate rises
above the equilibrium value, the core temperature rises, so the pressure rises. The core
therefore expands and cools, lowering the energy generation rate again. Hence the
equilibrium is restored (stable). The thermostat is very sensitive as dE/dt depends
strongly on temperature, e.g. p-p chain  2T4 and the CNO cycle 2T16
In degenerate stars, if the energy generation rises, the core temperature again rises, but
this has no effect on the degeneracy pressure. Thus there is no expansion or cooling.
However, the increased temperature will lead to more energy generation, so there is a
runaway thermonuclear explosion (e.g. the ‘helium flash’ at the onset of helium
burning in the degenerate cores of low mass post-main sequence stars).
Stable thermonuclear energy generation is therefore not possible in degenerate stars.
Page 8 in tranche 3 of 8 pages. There were 18 previous pages