Sunteți pe pagina 1din 33

Chapter 2

The standard solar model

We can get information of the internal structure of the Sun through three
main means: energy flux, neutrinos, and solar oscillations. Of these, the
energy flux is a strongly averaged source of information. Neutrinos are
difficult to observe but yield somewhat more specific information. They are
also the only way to get direct information from the solar interior. Solar
oscillations have become an increasingly important tool for detailed studies
of the interior of the Sun during the last 30 years. (We will return to
neutrinos and oscillations later).

2.1 Age and pre-main-sequence evolution

A useful solar model must yield correct luminosity, radius, and age of the
present Sun. In order to create the model we need to understand something
of the evolution history of the Sun. The reason is that we cannot directly
measure the helium content in the core, i.e., how much of its hydrogen
reservoirs the Sun has so far burned to helium.
As the hydrogen is fused to helium the mean molecular weight µ
increases. In order to keep the thermal pressure

P = ρRT /µ (2.1)

large enough for the Sun not to collapse the temperature and/or density of
the core must increase. (Here R is the gas constant, and ρ is mass density).
These changes enhance nuclear reactions, which implies an increasing energy
production and increasing luminosity. Thus our Sun is getting brighter,
slowly but unavoidably.
The best way to determine the age of the solar system is to study mete-
orites which are likely the oldest bodies of the solar system. An important

21
22 CHAPTER 2. THE STANDARD SOLAR MODEL

clue to their age is given by 87 Rb that decays to 87 Sr with a half-life of


4.8 × 1010 yr. Comparing the abundances of 87 Rb and 87 Sr relative to the
abundance of the stable 86 Sr the age of the solar system has been estimated
to be (4.55±0.05)×109 yr (according to recent studies the error margin may
be a factor of 10 smaller). Similar studies of thorium and uranium decaying
into lead isotopes confirm this estimate.
Hence the age of meteorites, i.e., the time elapsed since the condensation
and chemical differentiation of these bodies began, is well known. However,
we need to relate this epoch to the start of the Sun’s life in the main sequence.
For that, we need to know the duration of the various processes during the
formation of the Sun and solar system.
A plausible scenario for a formation of a Sun-like star is the following:

1. Take an interstellar gas cloud with a mass of the order of 104 m .


Some perturbation, e.g., an interstellar shock wave can lead to the
gravitational collapse of the cloud if its self-gravity exceeds the internal
pressure of the gas. This condition is called the Jeans criterion:
Gmc RT
> , (2.2)
r µ
where µ, T, mc , and r are the molecular weight, temperature, mass
and radius of the cloud.
In reality the collapse is not spherically symmetric due to rotation
and magnetic field. Angular momentum per unit mass for typical
interstellar clouds is 1018 m2 /s but for the present solar system only
1016 m2 /s .
Typical magnetic flux densities in interstellar clouds are 0.1 − 1 nT,
which, if compressed to the size of a solar system, would yield huge
intensities of about 106 T. Even the highest (natural) concentrations
of magnetic field in the present solar system are below 1 T.
Thus virtually all initial angular momentum and magnetic flux have
disappeared in our solar system. The magnetic field provides an ef-
fective method to remove the angular momentum. This phenomenon
is called magnetic braking. (It also plays a role in solar rotation
as we will see later.) Later, when the density has become sufficiently
large (but the temperature remaining still low enough), neutral atoms
are formed by recombination of electrons and ions, the connection of
matter (plasma) and magnetic field is lost and the magnetic field can
escape back into interstellar space.

2. During collapse the cloud fragments to form some 103 −104 stars. This
appears to happen during the so called free-fall time tf f which is the
2.2. BASIC EQUATIONS 23

time of a spherically symmetric cloud with negligible internal pressure


to collapse. A typical free-fall time is about 3 ×107 years.

3. Next the fragments collapse. This is possible because, when the density
increases, the Jeans criterion is satisfied for smaller masses (assuming
no significant increase in temperature). The center of a collapsing
fragment becomes optically thick and heats up until hydrostatic equi-
librium is reached. This is called a protostar.

4. The protostar evolves quickly (< 106 yr) to the main sequence of the
H-R diagram. A cool star is born. Its parameters are:
Tef f ≈ 3000 K
Tcore < 106 K (not yet hydrogen burning)
r ≈ 4 r
L > L

5. The cool star slowly contracts gravitationally. This heats its core until
hydrogen burning ignites. The Sun has then arrived at the zero-age
main sequence. The duration of the slow contraction period, the so
called Kelvin-Helmhotz time, is also about 3 ×107 years.
Accordingly, the present uncertainty in the dating of the meteorites is
less than the length of the whole pre-main-sequence evolution of the
Sun. The material from which the meteorites are formed condensed
during this pre-main-sequence evolution or soon thereafter. Thus the
uncertainty in the dating of the origin of meteorites with respect to
the Sun’s pre-main-sequence evolution is larger than the uncertainty
in the age of meteorites itself.
A reasonable conservative estimate for the (main sequence, hydrogen
burning) age of the Sun is

t = (4.57 ± 0.05) × 109 yr . (2.3)

During this period the luminosity and radius of the Sun have slowly
increased to their present values.

2.2 Basic equations

In this section we formulate a set of equations that govern the solar structure
and evolution.
24 CHAPTER 2. THE STANDARD SOLAR MODEL

2.2.1 Conservation laws

Conservation of mass

Conservation of mass (within a sphere of radius r) is usually given in the


form ∂m/∂r = 4πρr2 . However, in case of the Sun it is more convenient
to consider the mass m interior to a sphere of radius r as the independent
variable, whence the equation becomes
∂r 1
= . (2.4)
∂m 4πρr2

We assume that the Sun is almost in a steady state, but in general


r = r(m, t) and thus we retain the partial derivatives above. The reason to
take m as the independent variable is that, except at the very beginning,
the mass loss has been negligibly small and we know m throughout the
whole period to which our model will be applicable. Thus we can place the
boundary conditions at m = m . On the other hand, we do not know the
radius, except for the present Sun, and the model has to give it as a result.

Conservation of momentum

For most of the discussion we can consider the Sun to be in a hydrostatic


equilibrium, ∂P/∂r = −ρg. Using conservation of mass and g(r) = Gm/r2
we obtain
∂P Gm
=− . (2.5)
∂m 4πr4
Note that this equation can not describe the collapse of the protostar where
the time-dependent momentum equation (hydrodynamic balance)

∂2r ∂P Gm
2
= −4πr2 − 2 . (2.6)
∂t ∂m r
is needed.

Energy balance

Let L(m) be the luminosity generated inside the sphere of mass m, ε the
energy generation per unit mass, and S the entropy per unit mass (i.e., the
specific entropy). Then
∂L ∂S
=ε−T . (2.7)
∂m ∂t
During the main sequence evolution the interior of the Sun is very close to
thermal equilibrium and the heating/cooling effect (∂S/∂t) is small. The
function ε(ρ, T ) represents the nuclear energy sources to be discussed later.
2.2. BASIC EQUATIONS 25

2.2.2 Energy transport

The energy flux F is by definition the luminosity per unit area. The two
main forms of energy transport are by radiation (FR ) and convection (FC )

L
F = F R + FC = . (2.8)
4πr2
The relative importance of these two forms varies greatly with the depth of
the layer considered.

Radiative transport

Radiative transport dominates the solar energy transport in the radiative


zone out to the distance of about 0.7 r from the center of the Sun. In the
radiative zone photons carry the energy produced in the core. Energy is
transported outward because the emission probability of photons is greater
outward than inward due to spherical geometry.
The mean free path of photons is rather short in the radiative zone
because of continuous collisions with electrons and nuclei. Therefore the
energy transport is very slow compared to velocity of light. The average
energy of photons (which is related to temperature) decreases when moving
outward. Photons are in the gamma and X-ray range in the core and deeper
radiative zone, but in the EUV range in the upper radiative zone and in the
visible range on the surface.
This is described as follows. Let θ be the angle to the local vertical
direction, Iν the intensity, κν the absorption coefficient, and Sν the source
function. Then the radiative transfer is governed by the equation

dIν
cos θ = −κν ρ(Iν − Sν ) . (2.9)
dr
(Intensity is reduced between two subsequent layers by an amount which is
absorped and enhanced by an amount which is generated there).
In the interior of the Sun the photon mean free path is so small that
the place where the photon is emitted and absorbed have nearly the same
temperature. Thus the Sun is in local thermodynamic equilibrium
(LTE). This means that the distribution of atomic levels is described by the
Boltzmann distribution and the particle distributions are Maxwellian, all for
the same temperature T . Thus we can replace the source function by the
Kirchhoff-Planck function
2h ν3
Bν (T ) = . (2.10)
c2 ehν/kB T − 1
26 CHAPTER 2. THE STANDARD SOLAR MODEL

whence equation (2.9) becomes


dIν
cos θ = −κν ρ(Iν − Bν ) . (2.11)
dr
Note that we cannot write Iν (r) = Bν , because that would imply a per-
fectly isotropic radiation field with no net transport. Actually, energy is
transferred outward and we can expand the intensity as
cos θ dBν
Iν = Bν − , (2.12)
κν ρ dr

which satisfies equation (2.11) up to a term proportional to cos2 θ (which


does not contribute to Fν ).
The energy flux is
∞
F = Fν dν , (2.13)
0
where

4π dBν
Fν = cos θ Iν dΩ = −
3κν ρ dr
4π dBν dT
= − . (2.14)
3κν ρ dT dr
(All even powers of Iν vanish in integration).
Setting F = L/4πr2 and using (2.4) to replace dr by dm we obtain the
temperature gradient
∂T 3κL
=− , (2.15)
∂m 256π 2 σr4 T 3
where κ is the opacity (so called Rosseland mean absorption coefficient)
defined by
∞ 1 dBν

1 0 κν dT
= ∞ . (2.16)
κ dBν

0 dT
The opacity is a so called harmonic mean, i.e., its inverse is an average of
1/κν . Thus more energy is transported at frequencies where the matter is
more transparent. It is also a mean weighted by dBν /dT so that more energy
is transported at frequencies where the radiation field is more temperature-
dependent.
Aside of radiative transport, the other means of energy transport in the
radiative zone is the heat conduction which mainly proceeds via electrons.
Heat conduction by electrons forms a small fraction of energy transport in
the radiative zone.
2.2. BASIC EQUATIONS 27

Transport by convection

In the radiative zone In the upper radiative zone at about 0.7 r the tem-
perature cools down so low that more and more atoms can be formed by the
recombination of electrons and nuclei. This greatly increases opacity since
atoms can more effectively absorb radiation than free electrons and nuclei.
Accordingly, transport of energy by radiation becomes too inefficient and
the medium becomes unstable for convection.
The convection zone is like a layer of liquid or gas heated strongly from
below where warm liquid rises up, cools down and subsides down again.
These rising and subsiding flows form in the convection zone large convection
cells, so called super cells that are divided closer to the surface to smaller
and smaller cells.
This is seen on the solar (photospheric) surface as a granular structure.
The surface consists of small grains or granules of about 1000 km diameter
each. Hot gas is rising from below in the bright center of the granule and
cooler gas is subsiding in the darker regions between the granules. There is
also evidence for the existence of a meso-scale and large-scale granulation.
Convection is a very efficient means of energy transport and the temper-
ature gradient in the convection zone
 
∂T ∂T
= (2.17)
∂m ∂m C

is close to the adiabatic gradient (maximum gradient with no heat loss).


Convection will be discussed in more detail later.
NOTE: The energy transport equation is the last equation in the system
of four differential equations of the solar model (2.4, 2.5 or 2.6, 2.7, and 2.15
or 2.17) giving gradients of r, P, L, and T with respect to m.
In addition, so called ”constitutive” relations are needed for the remain-
ing four variables ρ, S, ε, and κ. These equations will be discussed below.

2.2.3 Equation of state

The pressure arises from momentum transfer by particles and photons. In


the case of the Sun the pressure has two components, the gas pressure (PG )
and the radiation pressure (PR ). The radiation pressure becomes important
only in the rarefied layers, the atmosphere and solar wind. Thus for the
present solar model we apply the ideal gas equation 2.1 for pressure.
The constituents of the gas are usually denoted by
28 CHAPTER 2. THE STANDARD SOLAR MODEL

X abundance (mass fraction) of H


Y abundance of He
Z abundance of heavier elements
Note that the ionization of elements adds particles to the gas and reduces
µ. E.g., the mean molecular weight of fully ionized electron-proton plasma
is half of neutral hydrogen gas.
Because Z is small, in the following discussion the heavy elements are
considered for simplicity to be fully ionized whence Z → Z/2. This intro-
duces only a small error. Then the mean molecular weight in the case of
completely neutral hydrogen and helium is
1
µ0 = . (2.18)
X + Y /4 + Z/2
Ionization of H and He adds particles reducing µ to
µ0
µ= , (2.19)
1+E
where E is the number of electrons set free by ionizing H and He divided by
the number of all other particles. E is given by
E = µ0 [ηH X + (ηHe + 2ηHe+ )Y /4] . (2.20)

The degrees of ionization ηH , ηHe , ηHe+ are given by the Saha equa-
tions
nH+ 2(2πme )3/2 (kB T )5/2
ηH = = exp(−χH /kB T )
nH u H h 3 Pe
nHe+ 2u + (2πme )3/2 (kB T )5/2
ηHe = = He exp(−χHe /kB T ) (2.21)
nHe uHe h3 Pe
nHe++ 2(2πme )3/2 (kB T )5/2
ηHe+ = = exp(−χHe+ /kB T ) ,
nHe+ uHe+ h3 Pe
where Pe is electron partial pressure, and the χ’s are the ionization potential
energies of H, He, and He+ . The u’s are partition functions of particles with
bound electrons 
ui = gij exp(−Eij /kB T ) (2.22)
j
where gij is the statistical weight of the jth state, and Eij is the energy of
that state, relative to the ground state.
For an isolated atom or ion (2.22) has an infinite number of terms and
diverges. However, most of the terms have energies close to the ionization
energy and cut off in a dense plasma where particles perturb each other
and lower the ionization potential. In practice the partition functions are
commonly approximated by the statistical weights of the ground state:
uH = 2 ; uHe = 1 ; uHe+ = 2 .
2.2. BASIC EQUATIONS 29

Figure 2.1: Ratio of electrostatic to thermal energy (upper curve) and elec-
trostatic pressure correction (lower curve).

Corrections to ideal gas law

In the ideal gas approximation the particles have only kinetic energy and
interact only by elastic collisions. However, charged particles also interact
by the electrostatic force and have related potential energy. The importance
of this correction to ideal gas law is estimated by the ratio of particles’ mean
electrostatic energy
EES = e2 /4π 0 r
to the average thermal energy 3kB T /2. In a good plasma this ratio is much
smaller than one. In the Sun it is on the order of ≤ 0.1. So it is not negligible
but small enough for the following Debye-Hückel treatment to be useful
as the leading correction to ideal gas law.
In the neighborhood of an ion the density of any other species with charge
eZ (Z = −1 for electrons) deviates from the mean density nZ  according
to the Boltzmann distribution
 
eZϕ
nZ = nZ  exp − , (2.23)
kB T
where the potential ϕ is determined by all charged particles. The potential
is found to be  
eZ −r
ϕ= exp , (2.24)
4π 0 r λD
30 CHAPTER 2. THE STANDARD SOLAR MODEL

where
 1/2
0 kB T
λD =  (2.25)
e2 Z 2 nZ 
is the Debye length. Expanding the potential to the first order we find the
electrostatic energy density

1 e3 ( Z 2 nZ )3/2
UES = eZnZ ϕES ≈ − < 0. (2.26)
2 8π 0 ( 0 kB T )1/2

Thus the electrostatic correction to the pressure PES = UES /3 is negative.


Thus, for given density and temperature, P < nkB T . The maximum cor-
rection is found at the depth where T ≈ 5 × 104 K and P ≈ 108 Pa, where
the correction is −5.9% (see Fig. 2.1). This happens somewhere half-way
up in the convection zone.
Another important correction is due to electron degeneracy. Electrons
are fermions and cannot be packed too closely together in the six-dimensional
phase space (r,p). Fermions follow Fermi-Dirac statistics which deviates
from the Maxwellian statistics at high densities. This correction becomes
significant at P ≈ 1014 Pa, i.e., inside r ≈ r /2 and reaches its maximum
of 1.7% in the center of the Sun (see Fig. 2.2).

Figure 2.2: Electron degeneracy in the solar core. Upper curve: degeneracy
parameter; lower curve: pressure correction.
2.2. BASIC EQUATIONS 31

2.2.4 Entropy

The second constitutive equation deals with entropy. Only the change of
the entropy dS is of interest, either in time or in depth. So we can use the
differential form common in thermodynamics but, instead of the extensive
variables U and V, we use the intensive variables P and T.
From elementary thermodynamics we have the relationship
 
dT dP
dS = cP − ∇a , (2.27)
T P
where  
∂S
cP = T (2.28)
∂T P
is the specific heat at constant pressure, and
 
∂ ln T
∇a = (2.29)
∂ ln P S

is the adiabatic temperature gradient (or, more precisely, the double loga-
rithmic isentropic temperature gradient).
Now we must determine cP and ∇a in terms of T and P . After some
calculation we find

∇a = , (2.30)
T ρcP
where  
∂ ln ρ
δ=− . (2.31)
∂ ln T P
To find cP we use the relation T dS = dU + P dV where U is the specific
internal energy and V = 1/ρ the specific volume (recall that ”specific” means
per unit mass). Then, with the help of the definition of Eq. 2.31 one can
prove that  
∂U Pδ
cP = + . (2.32)
∂T P ρT
Finally we need U in terms of P and T . Considering only the main contri-
butions, the kinetic energy and hydrogen and helium ionization energies we
have
3RT 1
U= + [nH+ χH + nHe+ χHe + nHe++ (χHe + χHe+ )] . (2.33)
2µ ρ

Now we can determine the specific heat and adiabatic temperature gra-
dient, insert them in the expression for dS and use this in the computations
of the solar model. Note that the deviations from the ideal gas expres-
sions cP = 5R/2µ and ∇a = 2/5 occur mainly in the layers where H and
He are partially ionized because there the degrees of ionization depend on
temperature (see Fig. 2.3).
32 CHAPTER 2. THE STANDARD SOLAR MODEL

Figure 2.3: Specific heat and adiabatic temperature gradient in the solar
interior as functions of pressure. The main excursion is due to the ionization
of H and first ionization of He, while the smaller hump is due to the second
ionization of He.

2.2.5 Energy production

Next we discuss the energy generation function ε defined in Eq. 2.7. From
the equation of state we know ρ(P, T ). Thus we can make a transformation
from ε(P, T ) to ε(ρ, T ). This is useful, as variables (ρ, T ) are more natural
in energy production calculations than (P, T ).
Solar energy is produced mostly by hydrogen burning to helium, where
6.683 MeV energy per nucleon is produced. One important point is that
the mass defect in helium is much larger than the mass defect of any other
nucleus compared to its building blocks. Therefore, the gained energy re-
lease is the largest of all nuclear reactions. Actually the whole fusion chain
involves 4 protons which finally fuse into one alpha particle, i.e., the total
release is 26.732 MeV.
Of all nuclei, protons have the smallest charge which is important to
get two particles sufficiently close to each other. In the case of protons an
electrostatic barrier of about 1 MeV must be overcome. This is significant
considering that the interior temperature of the Sun is about 1.5×107 K, i.e.,
2.2. BASIC EQUATIONS 33

1.3 keV only. Thus a very high density is required in order that sufficiently
many close encounters can take place.
There are two main reaction chains for nuclear fusion in stars. In the
Sun 98.8% of energy comes from the pp chain and about 1.2% from the
CNO cycle. In the discussion of nuclear reactions we adopt the notation
X(a,b)Y where
X is the target nucleus
a is/are the incident particle(s)
b is/are the emitted particle(s)
Y is the residual nucleus
Furthermore we denote p = 1 H, d = 2 H, α = 4 He etc., and let ∗ denote
an excited state. Note that tritium is not important in the Sun whereas
it is the main fuel component together with deuterium in present tokamak
experiments on energy production through controlled thermonuclear fusion.
In the following tables we label the various reactions by their reaction
rate symbols (λ). The total energy release per reaction is Q = Q + Qν ,
where Q denotes the energy delivered to the thermal bath (in this case:
charged particles) and Qν the (mean) energy carried away by the released
neutrino.

pp chain

The proton-proton chain has three branches


Reaction Rate symbol Q (MeV) Qν (MeV)

ppI p(p,e+ ν)d λpp 1.177 0.265


d(p,γ)3 He λpd 5.494
3 He(3 He,2p)α λ33 12.860

ppII 3 He(α, γ)7 Be λ34 1.586


7 Be(e− ,ν)7 Li λe7 0.049 0.815
7 Li(p,α)α λ17 17.346

ppIII 7 Be(p,γ)8 B λ17 0.137


8 B( ,e+ ν)8 Be∗ λ8 8.367 6.711
8 Be∗ ( ,α)α λ8 2.995
Of these ppII is an alternative to (λ33 ) and ppIII an alternative to (λe7 ).
The branching ratios depend on reaction rates, and therefore on tempera-
ture. For the present Sun the ratios are I:(II+III) = 85.2:14.8 and II:III =
14.8:0.019.
34 CHAPTER 2. THE STANDARD SOLAR MODEL

The last digits in this table depend on small details of the reaction co-
efficients used in model calculations. This table is from the 2nd edition of
Stix’ book and differs slightly from the same table in the 1st edition. The
most notable change is in the mean energy of the 8 B neutrino (6.711 MeV
instead of 7.2 MeV).
While the ppIII branch is of minor importance in the total energy pro-
duction it is the sequence which produces neutrinos that are easiest to detect
on the Earth in the large water and heavy water detectors due to their high
energy. (Neutrino observations will be discussed later in more detail).
Whatever branch the pp chain takes the total energy per produced α
particle is the same (26.732 MeV). Note also that in order to get one α
particle, two 3 He nuclei need to be produced in the two first ppI chain
reactions p(p,e+ ν)d and d(p,γ)3 He. Thus ppI destroys 4 protons in order
to create one α particle and, as a ”side-product”, 2 positrons, 2 neutrinos
and 2 photons.
Note that the first ppI reaction is a weak interaction process (mediated
by the heavy W boson), and therefore quite slow, occurring typically only
once in 1010 years. If it was a faster process, the Sun would have burned all
of its energy already long ago. Note also that neutrinos take a considerable
amount of the total released energy away from the Sun.

CNO cycle

In more massive stars whose cores are hotter than in the Sun, at least about
2×108 K, the so called CNO cycle dominates.
Reaction Rate symbol Q (MeV) Qν (MeV)

12 C(p,γ)13 N λp12 1.944


13 N( ,e+ ν)13 C λ13 1.513 0.707
13 C(p,γ)14 N λp13 7.551
14 N(p,γ)15 O λp14 7.297
15 O( ,e+ ν)15 N λ15 1.757 0.997
15 N(p,α)12 C λp15 4.966
Note that each of the heavy nuclei appear exactly once on both sides of
the nuclear reactions. Therefore, although they are transmuted in these
reactions into each other, their total amount remains the same. Thus, they
are not produced from lighter nuclei in these CNO chain reactions but only
act as catalysts in order to create one α particle from four protons, as in the
pp chain.
In the CNO cycle the same amount of energy, 26.732 MeV is released
per one α particle as in the pp chain. Also, as clearly seen in the above
2.2. BASIC EQUATIONS 35

Figure 2.4: Nuclear energy generation in the Sun. Energy production is


confined inside r < 2 × 108 m, i.e., one fourth of the radius, but only 1.5%
of the volume. Note that the 1.2% contribution by the CNO cycle is even
more concentrated within the core than the main pp cycle.

table, the same number and type of ”side-product” particles are produced
in either chain.
Note that the second and fifth CNO reaction are weak interaction pro-
cesses. However, they are faster than the first pp chain reaction. Therefore
the CNO cycle proceeds quite much faster than the pp chain. Actually, the
speed of the CNO cycle is set by the fourth reaction which, in solar condi-
tions, occurs typically once in 109 years, i.e., an order of magnitude faster
than the slowest process of the pp cycle. The nuclear rates, however, depend
greatly on temperature and in heavier stars with higher temperatures, the
rate of the CNO cycle may be much faster and even dominate the energy
production (and stellar evolution).
Note also that the neutrinos in the CNO cycle take away a larger fraction
of the total released energy than in the pp chain. Accordingly, the CNO cycle
is less effective in yielding energy to the thermal bath than the pp chain.
Now the remaining task is to calculate the energy production ε as func-
tion of distance from the center. This requires a careful quantum mechanical
calculation of the various reaction rates rik between each pair of species i
and k. Finally 
ε= Qik rik , (2.34)
36 CHAPTER 2. THE STANDARD SOLAR MODEL

where Qik denotes the thermal energy released in the respective reaction.
The calculation of the reaction rates and the ε function is too lengthy to be
discussed in these lectures. The result is shown in Fig. 2.4.

2.2.6 Opacity

As the final constitutive equation, also the opacity κ is natural to be given


in terms of variables (ρ, T ). Recall the definition (Eq. 2.16)

∞ 1 dBν

1 0 κν dT
= ∞ dBν
. (2.35)
κ

0 dT

Thus the first task is to determine the spectral absorption κν by calcu-


lating the absorption/scattering of photons by atoms, ions, and electrons.
This requires again lengthy time-dependent quantum mechanical calcula-
tions from which we get the cross sections σ for the various interactions.
Below we list the results for the most important contributions.

Bound-bound absorption

Bound-bound absorption simply means a transition of an atom or an ion to


a higher state of energy by absorption of a photon having the same energy
as energy difference between the two states. In analogy with a classical
oscillator, the cross section per atom is given by

e2
σbb (ν) = f φ(ν) , (2.36)
4 0 me c

where f is called the oscillator strength which contains the transition


probability, and φ(ν) is the normalized shape function of absorption or the
line profile expressing the frequency dependence. The two most important
line profiles are the Doppler broadening

1  
φD (ν) = √ exp −(ν/νD )2 (2.37)
πνD

which arises from the Maxwellian velocity distribution of the absorbing par-
ticles and has therefore Gaussian form, and the so called Lorentz profile
which is due to collision broadening
γ
φC (ν) = . (2.38)
(2πν)2 + γ 2 /4
2.2. BASIC EQUATIONS 37

In these equations ν = ν − ν0 is the distance from the line center ν0 , γ


is the constant of collisional damping that determines the collisional width,
and νD = (2RT /A)1/2 ν0 /c is the Doppler width (A is the atomic weight).
In general, both Doppler and collision broadening act together and the
two functions must be folded. The Doppler broadening dominates at the
core (or center) of the line close to ν0 and the collisional damping at the
”wings” of the line further out from ν0 .

Bound-free absorption

Bound-free absorption means absorption of a photon by ionization, i.e., pho-


toionization. The determination of the cross section requires complicated
analysis. For example, the cross section per a hydrogen-like atom or ion in
the state with principle quantum number n is given by

me e10 Z 4 gbf (n, ν)


σbf (ν) = √ (2.39)
48 3π 50 ch6 n5 ν 3

where gbf is the so-called Gaunt factor which arises from the quantum
mechanical calculation of the ionization probability and depends only weakly
on n and ν.
Obviously, for photoionization to occur, the energy of the photon hν
must exceed the ionization energy Uth,n which, for a hydrogen-like atom or
ion is
me e4 Z 2
Uth,n = (2.40)
8( 0 hn)2
For example the first ionization energy for H is 13.6 eV. This corresponds
to the frequency of 3.3 × 1015 Hz, or wavelength of 9.1 × 10−8 m = 91 nm,
which is in the EUV range.

Free-free absorption

A free electron can absorb a photon in the presence of a third particle which
can take the recoil momentum. This is an inverse process to bremsstrahlung.
The cross section

e6 Z 2 gf f (v, ν)
dσf f (v, ν) = √ dne (v) (2.41)
48 3π 2 30 chvm2e ν 3

depends on the electron distribution in the velocity space ne (v) and the
related Gaunt factor gf f .
38 CHAPTER 2. THE STANDARD SOLAR MODEL

If the small contribution from partial electron degeneracy is neglected,


the electrons can be assumed to be Maxwellian and the the total cross section
is simplified to

e6 Z 2 ne gf f
σf f (ν) = 3/2
(2.42)
24π 2 30 chme (6πkB T )1/2 ν 3

The free-free transition is possible for all energies. Thus σf f is a continuous


function of ν.

Scattering by free electrons

The scattering of photons from free electrons is called Thomson scatter-


ing. It can be treated classically as follows: The electron is set to oscillatory
motion by the incident wave electric field. The cross section can be calcu-
lated as the ratio of radiation emitted per unit time and incoming energy
flux density.
The classical cross section of Thomson scattering (per electron) is con-
stant
e4
σs = ≈ 6.65 × 10−29 m2 , (2.43)
6π 20 m2e c4

(independent, e.g., of photon energy), and is often given as

8π 2
σs = r (2.44)
3 0

where r0 = (e2 /4π 0 me c2 )1/2 ≈ 2.818 × 10−15 m is the classical electron


radius.
If the photon wavelength is longer than λD , collective effects affect the
electron scattering cross section to be reduced. Since electron scattering
is most important in the core, this reduces the opacity there by a few per
cent. This may look like a small factor but, because it reduces the central
temperature which affects the branching ratios of the pp cycle, this is a
critical effect when calculating the energetic neutrino flux from 8 B decay.

Calculating the Rosseland mean

Using all the above discussed cross sections it is finally possible to calculate
the radiative opacity κ per unit mass. The calculation is by no means simple.
The result as a function of pressure is given in Fig. 2.5.
2.3. SUMMARY OF THE MODEL 39

Figure 2.5: Opacity κ and its double-logarithmic temperature derivative


κT = (∂ ln κ/∂ ln T ) as functions of pressure.

2.3 Summary of the model

2.3.1 Differential equations and constitutive relations

Let us summarize the solar model discussed in previous sections. We have


four first order differential equations:

∂r 1
= (2.45)
∂m 4πρr2
∂P Gm
= − (2.46)
∂m 4πr4
∂L ∂S
= ε−T (2.47)
∂m ⎧ ∂t

⎪ 3κL

⎪ − in stable layer
∂T ⎨ 256π 2 σr 4 T 3
=   (2.48)
∂m ⎪
⎪ ∂T


⎩ in unstable layer
∂m C

In addition we have four constitutive relations

ρ = ρ(P, T ) (2.49)
dS = dS(P, T ) (2.50)
ε = ε(ρ, T ) (2.51)
κ = κ(ρ, T ) (2.52)
40 CHAPTER 2. THE STANDARD SOLAR MODEL

2.3.2 Boundary conditions

In order to integrate the four first-order differential equations we need four


boundary conditions. Two of them are convenient to be imposed in the
center of the Sun (m = 0): r(0) = 0 ; L(0) = 0.

First surface boundary condition

The other two boundary conditions are imposed on the surface (denoted by
index s). We first define the optical depth by
∞
τ (r) = κρ dr . (2.53)
r

We will show later that the effective solar temperature corresponds to the
temperature at the optical depth τ (rs ) = 2/3. Therefore we take this layer
to define the surface for the interior solar model. The solar atmosphere is
then the layer with r > rs , i.e., τ < 2/3.
We know the surface conditions rs , Ps , Ls , Ts for the present Sun only,
but we can derive two relationships for them, which are then the two re-
maining boundary conditions. In the whole atmosphere L = Ls (the energy
source is deep in the interior), r = rs (the atmosphere is geometrically thin),
and m = m (the atmosphere is very light).
In this situation using m as an independent variable would lead to very
inaccurate results. Therefore, we change the variable in the pressure equa-
tion from m to τ whence
∂P Gm
= 2 . (2.54)
∂τ rs κ
Integrating this through the atmosphere and assuming P (τ = 0) = 0 we get
the third (the first surface) boundary condition

2/3
Gm 1
Ps = dτ . (2.55)
rs2 κ
0

Second surface boundary condition

The second boundary condition is a bit more complicated. We assume first,


for simplicity, that the absorption coefficient κ is frequency-independent (so
called ”grey”). Then we integrate the equation of radiative transfer (see Eq.
2.11) over all frequencies to get it in the following form
dI
cos θ =I −B. (2.56)

2.3. SUMMARY OF THE MODEL 41

We use for the following expansion for the integrated intensity


∞
Iν dν ≡ I(τ, θ) = I0 (τ ) + cos θ I1 (τ ). (2.57)
0

This method is called the Eddington approximation. It is too coarse


for the study of the atmosphere itself but good enough for finding a proper
boundary condition for the model of solar interior. The functions I0 and
I1 can be determined by defining momenta in angular space

1
J = I dΩ
4π
F = I cos θ dΩ (2.58)

1
K = I cos2 θ dΩ .

The equation of transfer (2.56) then leads to the equations
dF
= 4π(J − B)

dK 1
= F. (2.59)
dτ 4π
Using the expansion of I 2.57 in the above definitions of J, F and K we
find J = 3K = I0 and F = 4πI1 /3.
Because F is by definition the energy flux which is constant in the surface
and above, we get J = B from the first equation of 2.59. Integration of the
second equation of 2.59 leads to
3
I0 = Fτ + b. (2.60)

Finally the constant of integration b is obtained from the radiation con-
dition that far from the Sun (i.e. at τ = 0) the net radiation into the upper
half space is the total flux F , whence b = F/2π.
On the other hand
I0 = B = σT 4 /π, (2.61)
and we find the fourth boundary condition
Ls
T4 = (3τ /4 + 1/2) . (2.62)
4πσrs2
This gives the atmospheric temperature as a function of optical depth. This
expression also shows that at τ = 2/3 the temperature is
Ls 1/4
T (τ = 2/3) = Tef f = ( ) . (2.63)
4πσrs2
42 CHAPTER 2. THE STANDARD SOLAR MODEL

Figure 2.6: The evolution of the Sun in Hertzsprung-Russell diagram.

2.3.3 Results of the model

Now we have a model which is generally called the standard solar model.
Of course, its details depend on the exact computations of the constitutive
relations and the assumption of the age of the Sun. It is clear that actual
computations are a very demanding task. Here we present only some of the
main results of the calculations.
Figure 2.6 shows the main sequence evolution of the Sun in the Hertz-
sprung-Russell diagram. The approach from the Hayashi line (describing the
non-nuclear hydrostatic evolution) to the start of hydrogen burning phase,
the so called zero-age main sequence (ZAMS) took about 5 ×107 years.
During its main sequence lifetime of 4.57 ×109 years the Sun has changed
very little in effective temperature and luminosity. This implies that evo-
lutionary effects can explain only a small part of the spread of of the stars
within the main sequence band.
Figure 2.7 presents the evolution of the solar radius and luminosity, and
the calculated neutrino counting rates for the 37 Cl and 71 Ga experiments.
(Solar neutrinos will be discussed later). The major part of the luminosity
increase is due to the increase in the solar radius from about 0.87 r to the
2.3. SUMMARY OF THE MODEL 43

Figure 2.7: Evolution of the solar radius and luminosity and the predicted
neutrino flux normalized to the present Sun.

present radius of 1 r . During the evolution the effective temperature has


only increased by about 125 K, thus contributing less to luminosity increase.
Figure 2.8 gives a table of the model results for the present Sun in nu-
merical form. The levels (lines) in the table are not equidistant in any of
the indicated variables; they are chosen so that both the outer layers and
the central region are reasonably resolved. The first line corresponds to the
surface at the optical depth of τ = 2/3.
Solar mass is indeed very concentrated towards the center. The outer
convection layer above about 0.71 r contains only about 2.5% of the total
mass. Only about 10% of mass lies outside 0.5 r in a region which makes
7/8 of the total solar volume.
Pressure and density change dramatically, by about 13 and 9 orders
of magnitude respectively, from the core to the surface. The change is
particularly rapid in the outer layers. Only in the highest layer above 0.9
r , pressure changes by about 8 orders of magnitude and density by about
5 orders of magnitude. In comparison, the variation of temperature is much
smaller, less than 4 orders of magnitude althrough and only 2 orders of
magnitude in the upper layer above 0.9 r .
Luminosity and hydrogen content vary only in the deep interior. No
luminosity is practically produced above 0.35 r . Hydrogen content remains
the same in the outer convection layer.
The mean molecular weight µ first decreases inwards from the surface
44 CHAPTER 2. THE STANDARD SOLAR MODEL

Figure 2.8: Some model calculated parameters at different altitudes of the


present Sun
2.3. SUMMARY OF THE MODEL 45

Figure 2.9: Mass fraction Y3 of 3 He in the present Sun.

due to the ionization of hydrogen and helium, but increases again near the
center where half of the hydrogen has already been converted into helium.
(The last column gives the polytropic (adiabatic) index Γ1 that is 5/3 except
just below the surface of the Sun.)
Due to the earlier mentioned temperature dependence of nuclear reac-
tions and the variation of temperature and density with the solar radius, the
fraction of the 3 He nucleus varies greatly in the different layers in the Sun.
The 3 He nucleus is destroyed quite rapily in the core but has a rather long
lifetime of about 109 years further out from the core at about one quarter
of solar radius. As seen in Figure 2.9 there is a dramatic concentration of
3 He nuclei in this layer whose peak is almost 100 times above the uniform

abundance of about Y30 = 4 ×10−5 .

Energy transport through the Sun

The energy produced by nuclear fusion is transported from the core by


gamma rays. Due to the high plasma density, the photons are continuously
absorbed and re-emitted by the gas in the radiative zone. The outward
energy diffusion is so slow that it takes, on an avearge, some 170 000 years
for a photon to reach the base of the convection zone at about 0.71 r . By
this distance the temperature has fallen from 15 million K in the core to
about 2 million K.
At the bottom of the convection zone, because of decreasing temperature
and the increasing formation of atoms from plasma, radiation becomes a less
efficient method to carry energy outward and transport by convection starts
to dominates above 0.75 r .
46 CHAPTER 2. THE STANDARD SOLAR MODEL

Convection is a much faster way of energy transport than radiation in


an opaque medium. It takes only about 10 days for the heated gas to rise
through the convection zone. During this process the gas also cools rapidly.
The solar surface is effectively a black body that absorbs all energy coming
from the convection zone and radiates it out at the temperature of 5778 K.

The future of the Sun

According to Figure 2.7 the radius and luminosity of the Sun have grown at
an almost constant rate. If the luminosity would now drop to 0.72 L , where
it was 4.5 billion years ago, the Earth would become ice-covered. Moreover,
increasing the luminosity thereafter back to the present level would never
thaw the ice because of the high albedo of an ice-covered planet.
However, geological evicende suggests that the whole Earth would ever
have been ice-covered. The solution to this ”faint-young-Sun paradox” most
likely lies in the evolution of the atmosphere. The greenhouse effect has
probably been more efficient than today due to larger amounts of ammonia
and/or carbon dioxide in the ancient atmosphere. That means that the early
atmosphere has been more efficient than presently in reflecting the escaping
infrared radiation back to the Earth. This could compensate the smaller
solar luminosity.
On the other hand, the Sun will keep on becoming brighter, slowly but
unavoidably, and this will have long-term consequences on Earth. Note that
this is a matter of hundreds of millions or billions of years, and should not be
confused with the ice-age cycles or with variations in solar magnetic activity.
Ice-ages are most likely related to the changes in the Earth’s orbital mo-
tion (eccentricity) and orientation (tilt angle, precession) whose time scales
are about 40 000–100 000 years.
The Sun has different quasi-periodic variations in its magnetic activity
which occur at short and long time scales from a few months to several
thousand years. The most well known of these is the 11-year solar cycle.
Solar irradiance variations probably also occur in relation with the varia-
tions in magnetic activity. Also, there is evidence for a correlation between
variations in solar magnetic activity and changes in global temperature on
Earth. E.g., the slow cooling trend during the last centennia of the previous
millenium which ended with a rapid warming during the second half of the
20th century greatly resembles the similar variation in solar activity. On top
of this, the recent ”global warming” is coincided by enhanced greenhouse
gas emissions due to fossil fuel burning. (We will discuss these issues later
in more detail).
In any case, in the very long run, the solar irradiation will increase greatly
according to the calculated solar evolution. At some point the atmospheric
2.4. SOLAR NEUTRINOS 47

temperature will have risen so much that the oceans begin to boil. There
is some controversy when this will happen as the different solar models give
different time scales for the last phases of hydrogen burning. Also, the
strength of future greenhouse effect is unknown. The best estimates vary
between 1 and 3 billion years.
So far the Sun has burned about half of its hydrogen content in the core.
After 5 billion years more it must begin to burn hydrogen in the outer layers.
At this time the Sun will leave its place in the main sequence of the H-R
diagram. At the age of about 12 billion years both the luminosity and the
radius of the Sun have increased by a factor of about 10 and the Sun has
become a red giant.
Thereafter the evolution is quite fast. After some 100–200 million more
years the red giant is assumed to flash for a while out to 100 r (i.e., beyond
the present orbit of Mercury) at the luminosity of about 1000 L . 100 million
years later the helium starts to burn to carbon in the core which causes
several flashes within the last 1–2 million years. Radiation pressure blows
the outer layers of the Sun into the interstellar space forming a planetary
nebula.
The hot inner core becomes a white dwarf some 12.3 × 109 years after
the birth of the Sun. The white dwarf is too light to compress further and
there will be no fusion of heavier elements. It will slowly cool down to lower
and lower temperatures.
In the case of sufficiently massive stars carbon will still burn to heavier
elements until 56 Ni and 56 Fe which have the largest nuclear binding energy
so energy is still released when they are formed. The formation of all other
elements heavier than these will need energy. Therefore they can only be
formed in non-equilibrium situations, like the burst of a (super)nova.

2.4 Solar neutrinos

One of the most famous problems in solar physics has been the so called
solar neutrino problem. It has long been one of the main reasons for further
refining the calculations of the standard solar models. In very simple terms
it means that the standard solar models predict a larger neutrino flux than
has been observed.
The above discussed pp-chain was suggested as the dominant solar en-
ergy production mechanism already in 1938. Before the first neutrino obser-
vations in 1967 the solar models had evolved so far that they could predict
with confidence the production of a copious amount of about 2 × 1038 neu-
trinos per second.
48 CHAPTER 2. THE STANDARD SOLAR MODEL

Figure 2.10: Predicted energy spectrum of the solar neutrino flux at 1 AU .


Units are neutrinos/(m2 s MeV) for the continua and neutrinos/(m2 s) for the
lines. Dotted vertical lines in the top mark the thresholds for the 71 Ga (233
keV) and 37 Cl (814 keV) experiments.

Figure 2.10 shows the calculated neutrino fluxes at 1 AU . The β-decay


channels produce continuous spectra, whereas the electron capture of 7 Be
produces two lines at 862 keV and 348 keV since the 7 Li nucleus can be
either in the ground state or in the first excited state. (Note that Qν = 815
keV given the earlier table is the weighted average of these two lines).
In addition to the reactions described above, also the so-called pep line
is shown in Figure 2.10. This is due to the reaction p(pe− ,ν)d (not included
in the ppI-III chain). Although this reaction occurs only at about 0.25% of
p(p,e+ ν)d, it is important because the neutrino has a relatively high energy
(1.442 MeV), clearly above the 37 Cl detector threshold. (Note also that in
this reaction practically all free energy goes away with the neutrino. Thus
it contributes to the Sun losing some energy). Only 8 B neutrinos (which
also come from a very rare reaction branch) have higher energy than the
pep neutrinos.
The energy of the solar neutrino is important, because the interaction
cross section increases linearly with energy. Therefore it is easier to detect
higher-energy neutrinos than the main (lower-energy) part of the neutrino
spectrum.
There are three basic types of solar neutrino detectors. The 71 Ga ex-
periment has the lowest energy threshold (233 keV), while the threshold of
the 37 Cl detector is 814 keV, and the large water detectors have the highest
2.4. SOLAR NEUTRINOS 49

threshold of about 5 MeV.


The first solar neutrino experiment started in the Homestake gold mine
in South Dakota in 1967. It used 615 tons of ordinary cleaning fluid, tetra-
chloroethene C2 Cl4 , whose chlorine is converted to argon in the reaction

νe + 37
Cl → e− + 37
Ar . (2.64)

The reaction also destroys the molecule but the radioactive argon atoms
remain in the vessel. The half-lifetime of 37 Ar is 35 days, which limits the
practical experiment length to about 100 days by which time the Ar atoms
must be counted at the latest.
The number of reactions is obtained by counting the argon atoms. This
is not an easy task as, on the average, one argon atom is produced in every
2.17 days. After about two months of observations 30 atoms are extracted
from among more than 1030 atoms altogether. This is being made with the
success rate of about 90% !
The next step was the Kamiokande water experiment in Kamioka, Japan,
that started operating in 1987. Water detectors detect neutrinos through
Čerenkov light from elastic ν − e− scattering if the recoil energy of the
electron is at least 5 MeV. While the water tank can only observe the 8 B
neutrinos, it is possible to determine their arrival direction and thus to know
how large a fraction of the neutrinos really comes from the Sun.
The lowest energy threshold is in the (inelastic) reaction

νe + 71
Ga → e− + 71
Ge , (2.65)

which produces a radioactive germanium nucleus (half-life of 11.4 days). In


1990 the Soviet-American Gallium Experiment (SAGE) began to operate
in Caucasus, and in 1991 the multinational GALLEX experiment in Gran
Sasso, Italy, using this reaction.
Due to the extremely weak interaction between neutrinos and ordinary
matter, a special unit for the observable solar neutrino capture rate has
been taken into use: the so called solar neutrino unit (snu). One snu
corresponds to the capture of one neutrino per second per 1036 target atoms.
Different model calculations give somewhat different capture rates for
the various reactions. Typical predictions for the gallium, chlorine, and
water detectors are
71 Ga: 130 ± 7 snu (GALLEX and SAGE)
37 Cl: 7.7 ± 1.1 snu (Homestake)
Water: 5.0 ± 0.75 ×106 cm−2 s−1 (Kamiokande)
Figure 2.11 depicts the predicted neutrino fluxes, cross sections and cap-
ture rates for the 37 Cl and 71 Ga experiments. Note how widely different the
50 CHAPTER 2. THE STANDARD SOLAR MODEL

Figure 2.11: Predicted neutrino fluxes, cross sections and capture rates for
the 37 Cl and 71 Ga experiments.

neutrino fluxes and (mostly in reverse order) the cross sections are for the
various neutrino sources, while the capture rates differ much less.
Figure 2.12 depicts the observed capture rates for the three types of neu-
trino experiments together with their predicted values. One can conclude
first that, no doubt, solar neutrinos have been detected. Second, all obser-
vations depict a smaller neutrino capture rate than predicted from the solar
theory. This is the famous solar neutrino problem that was realized first
in the chlorine experiments. The third conclusion is that the neutrino deficit
depends on neutrino energy. The chlorine detectors observe only about 2.5
snu, i.e., about one third of the predicted flux. The gallium detectors have
the best record, some 50% of the predicted value, and the water detectors
fall between these two.
The neutrino deficit has been attributed to several different causes. In
principle, it could be an observation problem. This was a popular explana-
tion as long as the Homestake observations were alone. But after the gallium
and water detector results, this appears very unlikely. For a long time it was
thought that the origin of the problem would be in erroneous solar models.
Quite a number of attempts have been made to correct the models but with
little success. Too simple or drastic fixes easily lead to problems somewhere
else in the models or with other observations.
One strategy to look for non-standard solar models has been to re-
duce the temperature, which would reduce the 8 B neutrino flux and help
at least with the original chlorine experiment problem. This is a reasonable
idea, as the 8 B neutrino rate is proportional to Tc18 , where Tc is the temper-
ature in the center. Lowering the central temperature by about 6% would
be sufficient to solve the 8 B neutrino deficit problem. However, this would
not be quite enough to decrease the 7 Be neutrino flux whose temperature
dependence is only ∝ Tc8 . Furthermore the lower-temperature solar models
would meet problems also with other observations, in particular with solar
oscillations.
2.4. SOLAR NEUTRINOS 51

Figure 2.12: Observed and predicted neutrino capture rates for the 37 Cl,

water and 71 Ga experiments.

Another attempt within non-standard solar models has been to reduce


the relative abundance of heavy elements (low Z). Since heavy elements are
good absorbers, this would reduce κ and thus the temperature gradient and
the core temperature. However, this model also leads to a rather low value
for the helium fraction, even lower than the primordial (big bang) value.
Morover, the solar oscillation observations contradict with this suggestion
as well.
Another non-standard solar suggestion was a rapidly rotating core which
would allow to lower the thermal pressure which is proportional to temper-
ature. However, a sufficiently large effect would need a very fast rotating
core (some 500 times faster than surface) that would lead to an oblate core
and to a gravitational quadrupole moment. This model is also inconsistent
with oscillation results that indicate that the core rotates at nearly (but not
exactly) the same angular speed as the surface.
Furthermore, a strong internal magnetic field could lower the central
temperature. However, the field should be very intense in order to have a
significant, say 10%, contribution to the pressure of the core. The origin of
the field would need to be in the primordial cloud but its life-time against
Ohmic dissipation would be much less than the age of the Sun.
Since recently it is evident that the solution really is in the physics of
neutrinos. The neutrinos appear in three different flavors (νe , νµ , ντ ), one
for each family of elementary particles (2 quarks and 2 leptons). Certain
52 CHAPTER 2. THE STANDARD SOLAR MODEL

non-standard elementary particle models predict that the neutrinos have


finite masses. In such a case the mass eigenstates are a superposition of the
three neutrino flavours and the neutrinos can oscillate from one flavor to
another. The nuclear reactions in the Sun produce electron neutrinos (νe )
only and the predicted fluxes are for electron neutrinos. If a large enough
fraction of solar neutrinos would transform to other neutrino flavors before
reaching the Earth, this would be an elegant solution to the whole problem.
In 1998 the Superkamiokande observations indicated oscillations between
muon neutrinos (νµ ), produced by cosmic rays in the atmosphere of the
Earth, and tau neutrinos (ντ ). This did not solve the problem of solar
neutrinos directly but it gave strong evidence that neutrinos are not massless
particles.
In the summer of 2001 the first results from the new heavy water (D2 O)
detector at the Sudbury Neutrino Observatory (SNO) indicated that the
solar neutrino problem was coming to its final solution. The point is that
this 8-kiloton water detector with 1 kiloton heavy water at its heart is able
to distinguish between the total neutrino flux and the νe flux.
Ordinary water detectors detect neutrinos through Čerenkov light from
elastic ν − e− scattering if the recoil energy of the electron is at least 5
MeV. All neutrino flavors can scatter in such a way but the combined cross
section of νµ −e− and ντ −e− reactions is a factor of 7 smaller than the cross
section of νe − e− scattering. Thus in the ordinary water detectors the other
neutrino flavors are practically hidden in the observational error margin.
However, in heavy water the deuteron can be broken in two different ways
by neutrinos. The first is the quasi-elastic process

ν + d → ν + p + n, (2.66)

where the neutrino can be of any flavor. This leaves a neutron whose sub-
sequent capture can be recorded by the release of an identifiable γ. The
second process is inelastic

νe + d → e − + p + p . (2.67)

and involves νe only. This is an inverse-β-decay which can again be observed


through Čerenkov radiation.
The present status of neutrino experiments is that the neutrinos indeed
oscillate between the different flavor states. The νe ’s from the Sun oscillate
at a high rate into νµ and less into ντ before reaching the Earth. Note also
that (solar) matter also has an effect on neutrino oscillations (the so called
Mikheyev-Smirnov-Wolfenstein effect) which must be taken into account.
The very good consistency of the standard solar model with solar os-
cillation observations, the failure of the non-standard solar models to solve
2.4. SOLAR NEUTRINOS 53

the neutrino problem and the new neutrino observations indicate that “the
solar neutrino problem” has been solved and the solution lies in the prop-
erties of the neutrinos. This is an excellent illustration of the strength of
the physical method. Not only is the answer in our reach but also the failed
attempts to find the solution within the non-standard solar models have con-
tributed enormously to our detailed knowledge of the interior of the Sun.
Had the neutrino deficit problem never existed, we would most likely have
been content with much less detailed models of the Sun itself.
Note still the different change of 37 Cl and 71 Ga neutrino fluxes during
the solar history, depicted in Figure 2.7. This difference can be attributed to
the different temperature dependence of the neutrinos observable by these
two experiments. As mentioned above, the 37 Cl neutrino flux comes from
the ppIII branch and is extremely sensitive to temperature. When the core
temperature increases from 1.35 ×107 K to 1.56 ×107 K during the solar
lifetime so far, the 37 Cl neutrino capture rate increases by more than an
order of magnitude. On the other hand, the lower energy threshold of the
71 Ga experiment allows the capture of a substantial fraction of neutorinos

from the ppI branch. Since the same branch is also mainly responsible for
energy production, the 71 Ga neutrino flux follows quite closely to the change
of luminosity and has only doubled during the solar lifetime.

S-ar putea să vă placă și