Está en la página 1de 14

Branden T.

Allen
April 24, 2001
Polytropic Star Models
During the 19
th
century many advances in many different fields of physics took
place, creating a new atmosphere conducive to the continuing investigations into the
natural world. One such field, that was investigated rigorously was that of the stars.
Through out the history of mankind, and many years before, it was evident that the sun had
not changed its behavior much. This is a fact that led to the original idea that one could
consider a star a gas sphere, that essentially must remain in hydrostatic equilibrium with its
own gravitation. Any situation contrary to this would lead to accelerations either inward or
outward, which would cause many short term changes, that have not been observed in
normal stars over the years.
1

Although this is an amazing achievement in ones consideration of the stars, a full
understanding of the complex processes was still not at hand. This description, though
informative, still does not address the question of why stars radiate, or what a stars energy
source may be. Half of this question was answered with the introduction of quantum
theory of blackbody radiation by Max Plank in 1900.
2
After the introduction of the
blackbody spectrum, it was determined that a star was essentially radiating according to the
rules of a black body, with the notable exception of the Fraunhofer absorption lines, which
were discovered by Fraunhofer in 1819, and emissions lines, which originate from the
stellar atmosphere and surrounding gas. This, however, still did not explain the stars
energy source. Once it was suggested that a star may live from its gravitational potential
energy, but this is not the case, as such a star would have only a fraction of the lifetime of a
real star. This being the case a new assumption was made, that the star had some sort of
undefined internal energy source. With this in mind Karl Schwarzschild began his work on
radiative transfer of energy in stellar atmospheres around 1906.
3
This established the
addition of a new dynamic to stellar structure, whose importance is equal to that of the first
assumption based on hydrostatic equilibrium. It was essentially determined that the
structure of a star was supported by internal pressure as well as radiation pressure against
its gravitational pull.
With this theoretical basis in place another important milestone was reached at
about the same time. Hertzsprung and Russell recognized that all stars are not of the same
mass, temperature, and luminosity and therefore have a number of differing properties
dependant upon these quantities. In measuring the luminosity and effective temperature of
each star they found that stars of the same class occupy distinguishable places in relation to
these quantities. With the creation of the Hertzsprung-Russell diagram many of the
classes, such as the main sequence stars and red giants, that are quite familiar today first
became apparent. Soon after this connection between star class and effective temperature
was established, Eddington established a mass-luminosity relationship for the main
sequence stars and further researched the internal mechanisms of the stars. In the 1930s
the Dirac Fermi statistics for a degenerate electron gas were published. It was quickly
realized that this too was applicable to the stellar interiors of certain star classes,
specifically white dwarfs of high mass. S. Chandrasekar was instrumental in this
determination, moreover he was also able to find the critical maximum mass for a white
dwarf star, known today as the Chandrasekar mass.
It is at this point in history where the Polytrope representation of stars first was
developed. It is a method that today still lends valuable methods and insights to the
1
Branden T. Allen
April 24, 2001
internal structure of stars. It has also proven to be most versatile in the examination of a
variety of situations, including the analysis of isothermal cores, convective stellar interiors,
and fully degenerate stellar configurations. Even the case of an ideal gas can be related to a
polytrope of index n = 3/2. The justification for such a theory is that, as the name implies,
it is extremely versatile. As will be shown later such a class of models allows for the
derivation and prediction of many stellar properties, which continue to be of significant
interest to the astrophysics community. The derivation of polytropic star models according
to R. Kippenhahn and W. Weigert, as well as S. Chandrasekar and William K. Rose is
outlined in the following section.
The polytropic theory of stars essentially follows out of thermodynamic
considerations, that deal with the issue of energy transport, through the transfer of material
between different levels of the star. We simply begin with the Poisson equation and the
condition for hydrostatic equilibrium:
Eq. 1 & 2
where P is the pressure, M(r) is the mass of a star at a certain radius r, and is the density,
at a distance r from the center of a spherical star.
4
Combination of these equations yields
the following equation, which as should be noted, is an equivalent form of the Poisson
Equation.
Eq. 3
From these equations one can then obtain the Lane-Emden equation through the simple
supposition that the density is simply related to the density, while remaining independent of
the temperature. We already know that in the case of a degenerate electron gas that the
pressure and density are ~ P
3/5
.
5
Assuming that such a relation exists for other states of
the star we are led to consider a relation of the following form:
2


P
r
M G
r
M
r
r
r
r

2
2
4
1
4
2
2
r
d
dr
r dP
dr
G

_
,

P K
n

1
1
Branden T. Allen
April 24, 2001
Eq.3
where K and n are constants, at this point it is important to note that n is the polytropic
index. Using this as a basis to classify different interior states within the star we can also
conclude that a non-relativistic degenerate electron gas is a Polytrope of n = 3/2. Based
upon these assumptions we can insert this relation into our first equation for the hydrostatic
equilibrium condition and from this rewrite equation to:
Eq.5
Where the additional alteration to the expression for density has been inserted with
representing the central density of the star and that of a related dimensionless quantity
that are both related to through the following relation.
Eq.6
Additionally, if place this result into the Poisson equation, we obtain a differential equation
for the mass, with a dependance upon the polytropic index n. Though the differential
equation is seemingly difficult to solve, this problem can be partially alleviated by the
introduction of an additional dimensionless variable , given by the following:
Eq. 7
Inserting these relations into our previous relations we obtain the famous form of the Lane-
Emden equation, given below:
Eq.8
3
K n
G r
d
dr
r
d
dr
n
n
( ) +

1
]
1

_
,

1
4
1
1
1
2
2


n
r a
a
n K
G
n

+
'


( ) 1
4
1
1
1
2
1
2
2
r
d
d
d
d
n

_
,

Branden T. Allen
April 24, 2001
At this point it is also important to introduce the boundary conditions, which are based
upon the following boundary conditions for hydrostatic equilibrium, and normalization
considerations of the newly introduced quantities and . What follows for r = 0 is
Eq.9
Eq. 10
Taking these simple relations into consideration, it is also evident that one can produce
additional conditions, based upon a modified form of the Lane-Emden Equation given by:
Eq. 11
Here it is apparent that as approaches 0 the first term of the equation approaches . As a
result an additional condition must be introduced in order to maintain the conditions of Eq.
9 and 10 simultaneously:
Eq. 12
Once the boundary conditions have been determined it is an easy matter to obtain a number
of solutions for the Lane-Emden equation. In addition to various numerical methods,
which will be explained later, this equation actually has 3 known analytical solutions for
polytropes of index n = 0, 1, and 5, given below:
Eq. 13, 14 & 15
4
( ) 0 1
r 0 0
2
2
2


d
d
d
d
n
+
d
d

0
0

0
2
1
1 1
5
2
1
6
1 6
1
1
1
3
+

+

,
sin( )
,
,
Branden T. Allen
April 24, 2001
where the subscripts represent the index number n for a specified solution and the
subscripts represent its value for = 0.
These results are useful in a few respects and deal with some actual state equation
for stars, however they are more important for inferring general forms of the Lane-Emden
solutions. Below are the plots of vs. for the aforementioned solutions, generated using
Mathmatica.
Polytrope n = 1
Please note that all the following graphs are modeled after the following: x axis gives
values and the y axis give values. Here we can see that the function basically follows
the same form as that for an index n = 0, with a few minor differences, however the
Polytrope of index n = 0 also terminates at a finite radius just as is observed in the relation
for a Polytrope of index n = 1. The other main difference that we observe in this case is
that the termination point of the star is markedly larger, at about a value of 3,15 as opposed
to the termination value of about 2,48 for the polytrope n = 0 as seen below.
Polytrope Index n = 0
5
Branden T. Allen
April 24, 2001
Here we can see that the values for start at one in accordance with our boundary
conditions and then eventually reduces to = 0. This is essentially an indicator that the
stars material ceases to exist outside of this area as the density drops to 0 at this point. The
next graph shows many similarities to this one.
Polytrope Index n = 5
Though these two solutions for n = 1 and n = 0 share many characteristics, the solution for
the Polytrope of index n = 5, contains some radically different and unexpected
characteristics. In this case the behavior of the function is markedly different than that of
its predecessors. Here the density of the star initially decreases rapidly as radius increases
but slows rapidly once a value of around three is reached. At this point the decrease
slows continually. Though it may not be apparent on the graphic provided, the function
never reaches 0. It is therefore evident that a polytropic star of index n = 5 has an infinite
radius, and in reality cannot exist, however the case itself deserves further study. Despite
this fact such a model provides important theoretical perspective concerning the theory, as
one may view this as the border between polytropic that are physically feasible. It is also
of interest to note, as will be shown later, that such a stellar model has, in spite of the
infinite radius, a finite mass. Additionally, other stellar models, that are created in a
layered fashion, where each layer consists of a Polytrope of a different index, may also
utilize this function for a portion of the star, in which case a finite radius would be possible.
In addition to these relations there are also a number of other conclusions that one
can draw from the polytropic model of stars. For relations of this type, there exists a
relation between the polytropic index, mass of a star and the radius. It is perhaps evident in
the discussion of the analytic solutions of the polytropic index that one could possibly infer
a relation between the polytropic index of the star and the radius that one would calculate
from that star. In the attempt to find a relation the most immediate result is obtained from
the simple equations of stellar state. Integration of dM/dr on the limit from 0 to the radius
of the star, yields the following expression.
6
Branden T. Allen
April 24, 2001
Eq. 16
This equation, is in itself not possible to integrate as we have an r dependence in the
density . However, this problem can be easily alieveated through the use of the polytropic
variables. By replacing and r by , where lambda is the central density, and a the
process of integration can not only be carried out more easily, but a relation dependent
upon central density the maximum Radius and the other polytropic variables obtained.
Integration after such a substition yeilds the follwoing:
Eq.17
Though this relation is extremely useful in the polytropic model of stars there are also a
number of additional relations, that can augment usefulness of this relation as well. The
first of these is the relationship between the central density and the average density of the
star. It is easily seen, that for the case of a star the average density is simply given by the
division of the total mass through the total volume. For the assumption of a spherical star,
which yeilds a volume of 4/3r^3, and using the relation for the total mass. We find simply
that the ratio of the average density to the central density is given by:
Eq.18
Where is the density and <> is the average density.
Another case, though not nearly as important as the case presented above, is found
in the relationship for the radius of a star. Beginning with the standard polytropic for the
radius (Eq.6) relation and inserting the value for a we obtain:
Eq. 20
7
M r dr
tot
R


4
2
0

M R
d
d
tot

1
]
1

4
1
3
1
1


< >



M
V
d
d
tot
3
1
1
R
K n
G
n
n

1
]
1

( ) 1
4
1
2
1
2
1


Branden T. Allen
April 24, 2001
It is now apparent that the case of n = 1 will, as in the case for the total mass of the star at n
= 3 which will be covered shortly, yield an expression that is independent of the central
density of the star. This relation is extremely interesting, as it reveals that there is the
potential for fundamentally different classes of stars to exist, whose masses differ, but not
the radii. Though this may seem counterintuitive it is important to note that as the mass in
a real star increases, the equations of state may change, which limits the actual range of the
masses for such a star in practice. This, however, does not diminish from the theoretically
interesting aspects of this model.
There is additionally another case, which is also independent of the central density
as well. For the case of a Polytrope of index n = 5, an analytic solution is obtainable, as
was shown above. Insertion of this relation for into the equation for total mass yields a
relation that in itself appears rather mundane, however we may also insert the value for the
maximum radius, which has already been determined to be infinite. By evaluation of the
resulting limit:
Eq. 21
It is readily apparent that the total mass of this star with an infinite radius yields the
surprising result of a finite mass. Though the significance of this case is in practice not
important, as there are not any stars that exist with an infinite radius, it is a case of
theoretical interest, as it essentially represents a border case in polytropic star models, as
models with n > 5 diverge. This case, however, is not nearly groundbreaking or interesting
as that of the first relation or the one to be discussed next, it is merely a theoretical model
that provides an interesting point of consideration, though composite models may contain
such polytropes and make use of such a relation.
In addition to the relations that we discussed here there remains an additional
relation between the mass and the radius that remains to be discussed. In this case we start
with the equations for the maximum radius of the star as well as the total mass for a star of
polytropic index n. Between these two equations, we can eliminate the central density and
obtain a relationship that is independent of this quantity. Through this we obtain the result
given below:
Eq. 21, 22, & 22
8
M a
d
d
a
K
G
tot

'

1
]
1

4 4 3
3
1
3
1
2
2
5
1



N GM R K
N
n w
w
d
d
n
n
n
n
n
n
n
n
n
n
n
n

1
]
1

1 3
0
1
1
0 1
1
1
1
1
4
1


Branden T. Allen
April 24, 2001
This is one of the most important results of the polytropic theory of stars, as it not only
provides, for a given radius a mass or vice versa, but also provides, in its simplified form,
some of the framework that is used in order to obtain numerical results for the other
intrinsically important values for a star. These variables and notation are usually used
rather universally in the literature regarding polytropic star models. An additional point
that is of paramount importance is the fact that a polytrope of index n = 3 will yield a mass
expression independent of the radius. As earlier discussed the case where the n = 3
corresponds to the polytropic star representation of a relativistic degenerate electron gas.
This is essentially the model for the Chandrasekar mass for stars (given as =1.54 x Suns
mass), and this can indeed be obtained by inserting value obtained from the polytropic
models in combination with the state equation of a relativistic degenerate gas. Though this
may not seem important, as the Chandrasekar border mass is a well known value, it is
important to point out that this derivation for the mass is a result as it is in complete
agreement with a value that was originally calculated through conventional methods
through the use of statistical mechanics. This argument not only lends credence and further
support to the border mass concept, but it also lends credibility to the polytropic theory of
stars, while demonstrating the effectiveness of the method. This is further bolstered by the
fact that there are no known white dwarfs that exceed this mass. This is a powerful
statement with regard to the theory of stellar states, that has been widely accepted for a
number of years.
This being said there remains one final relation for the central pressure that is easily
derivable from the expression for the central density. In order to obtain this relation we
begin with the expression for the central density given by,
Eq. 25
As this relation is dependant upon K and the central density it is apparent that we can
replace both of these relations by the values determined for the Mass Radius relation as
that determined for the value of the average density in relation to that of the central density.
The average density can then be replaced by its dependence on the mass and the volume
and these relations finally yield:
Eq. 26
9
P K P K
n
n
c
n
n

+ +

1 1
P N GM R w
GM
R
w
n
d
d
c n
n
n
n
n
n
n
n
n

1
]
1

1
]
1

'

1 3 1 2
4
2
1
4 1
1


( )
Branden T. Allen
April 24, 2001
Though it is often not mentioned it is interesting to note that this relation can also be used
as a good check on the accuracy of ones stellar model. It is apparent that as the central
pressure changes that the equation of state may change with it as well. For example if it is
known that the electron gas is degenerate we have two cases for the degeneracy. The first
of these is the relativistic case and the second of these is the non relativistic case. Each of
these possesses its own equation of state that is different the first. For the relativistic case
we have a polytrope of n = 3 and for the non - relativistic case we have a polytrope of n =
3/2. In any case if we have a star of a known mass, radius, and general chemical
composition one could easily determine the polytropic index. For the case of n = 3 this is
of extreme importance as we are able to determine the border mass. This gives a maximum
value of the central density before the collapse of a star into a neutron star, and the value
for the non-relativistic case give values that correspond to approaching values of largest
central pressure. At the point where the model exceeds the pressure of the Chandrasekar
mass we know that the model most certainly is that of a neutron star or just invalid. In any
case it provides a nice guide to the models of the stars that one would be concerned with
and also provides a good basis for the inclusion of models or exclusion of models, based on
the validity of the equations of state at different values for the pressure.
Having demonstrated the importance of the polytropic star relations, it is now
important to consider cases, in which one is not able to use the exact solutions of the
polytropic relations. Examples of such systems would be that of a non-relativistic electron
gas, where, as mentioned earlier, the polytropic index is n = 3/2. Even the all important
case of the fully degenerate electron gas (n = 3) is not covered by the exact solutions.
There are a number of ways that this problem could be attacked. The first of these is to use
a program such as Mathmatica in order to obtain solutions to the Lane Emden
equations. However under the assumption that one does not have access to such a program
there are also a number of other ways that the problem at hand can be attacked. The first of
these methods is to attempt a power series solution for the Lane-Emden equation:
Eq. 11
One can then proceed to formulate a power series solution to the Lane Emden equation.
Which is of the general form:
Eq. 31
where c subscript n represent arbitrary constants.
By matching the coefficients and inserting the appropriate border conditions, the resulting
solution is given by
10
2
2
2


d
d
d
d
n
+
( )

c
n
n
Branden T. Allen
April 24, 2001
Eq. 32
One can see immediately that this result is consistent with the analytical solution of n = 0,
as the additional terms for the series solution are n dependant. These solutions, though not
the only available for use in determination of physical quantities, are generally satisfactory
for representing different polytropic indices.
A method that is readily available is a variation of the Euler method, for second
order differential equations. In considering this method on creates an arbitrary step,
suitable to the task at hand, and from initial values of the equation extrapolates a solution.
In this case one considers the initial value of d/d as well as . At different itterations
one can now extrapolate values from the functions slope. This yields a recursive relation
for this method of:
Eq. 33
where h is the step size and the primes represent derivatives of the function . Using the
initial values and the Lane-Emden equation, one can then find approximate values of the
function and for the next step. Reinserting these values yields a new value of
allowing one to extrapolate another value. Repetition of this process allows one to crate an
approximate data curve, thereby describing a polytrope of a certain type.
Although this method may be of use it is quite evident that there remain a number
of other techniques that may be superior, which result in numerical results that one can use
in order to determine the necessary factors that may be placed in the various relations,
which were introduced in the preceding section, in order to make predictions for different
stellar properties. The first, and also the most used throughout the literature (See
Schwarzschild, Chandrasekar, Kippenhahn and Weigert) is the Runge-Kutta Method. In
order to determine the numerical solutions of these equations, one must first begin with
initial values as in the preceding section. The first of these is given by our normalization
condition, and the second of these is acquired through observation of the Lane-Emden
equation, in the expanded form, given by equation 13. As the normalization conditions
requires that (0)=1 it is apparent that the second term of the equation will diverge if the
value of d(0)/d is not equal to zero. It is therefore apparent that a second condition
exists, which makes the determination of solutions through numerical methods possible.
The general procedure for the Runge-Kutta method is as follows. As the Lane-Emden
equation is a second order differential, one must use an expanded form. According to the
methods outlined in a number of mathematical texts the various values of , d/d, for
11


( ) ... + 1
1
6 120
2 4
n


' ' ' '
' '
i i i
i i i
h
h
+
+
+
+
1
1
Branden T. Allen
April 24, 2001
certain values of can be determined through a step mehtod. The size of the steps h are
taken to be constant and provide an iterated view of the function ().
6

Eq. 31
Where the separate coefficients k are given by,
Eq. 32
Calculation of these parameters is in this case not limited to the exact polytropic solutions,
but can be tabulated for any polytropic index n, note that n of the Runge-Kutta method is
not n the polytropic index. As an example a list of a number of solutions is included to
below.
Where each of these variables represent the variables defined in the primary relations for
polytropic relations.
7

Now in order to convince the reader of the validity of such relations, in terms of
actual theory data agreement, the model of a fully degenerate white dwarf is considered.
As indicated earlier such a model is that of a polytrope with index n = 3. It is also of use to
remind the reader that the mass relation introduced earlier produces the Chandrasekar
mass, which defines the maximum mass of an exhausted star, before which it would simply
collapse further until it would reach the state of a neutron star. Inserting the appropriate
values into the mass radius relations, given by Equations 21, 22, and 23, we find the
radius of the star to be on the order of 28km - 40km.
4
This radius is necessarily much
12
n ? ?(d?/d?) ?(central)/?(average) o_w_n N_n w_n - [(n+1) ? [ d?/d? ] ]^-1
0 2.4494 4.8988 1 0.333 - 0.11936 0.5
1 3.14159 3.14159 3.28987 - 0.63662 0.26227 0.5
1.5 1.65375 5.99071 5.99071 132.3843 0.42422 0.7714 0.53849
3 6.89485 2.01824 54.1825 2.01824 0.36394 11.05666 0.85432
4 14.97755 1.79723 622.408 0.729202 0.4772 247.558 1.66606
5 Infinite 1.73205 Infinite 0 Infinite Infinite Infinite
( )

n n n n n n
n n n
h
k k k k
h
+
+
+ + + +
+
1 1 2 3 4
1
6
2 2 ' '
'




' ' ' ( , , ' )
' '
( . , . ' , ' . )
( . , . ' , ' . )
( , ' , ' . )
n n n n
n n
n n n n n n
n n n n n n
n n n n n n
f
k
k f h h k
k f h h k
k f h h k

+ + +
+ + +
+ + +
1
2 1
3 2
2 3
5 5 5
5 5 5
5
Branden T. Allen
April 24, 2001
smaller than that of a typical star, but larger than that of a neutron star, on the order of
~10km. This being a crude, however satisfying, example of a polytropic star models
practical use, is only one of many conceivable uses for such a theory.
Throughout this survey on the Polytropic theory of stars, it has been demonstrated
that in many respects these models provide a versatile and accurate method for the
determination of many stellar parameters. Though the Lane-Emden equation itself only
has three known solutions, this difficulty can easily be overcome with the introduction of a
number of numerical solutions, that can be found through a numerous variety of methods,
for which approximate solutions can be found to varying degrees of accuracy. This being
the case the polytropic theory of Stars has proven to be a versatile and able theory worthy
of its name.

13
1
Martin Schwarzschild. Structure and Evolution of Stars.
Princeton University Press, Princeton 1958
22
H.H. Voigt. Abri der Astronomie (5. berarbeitete Auflage).
Bibliografisches Institut und F.A. Brockhaus A.G., Mannheim 1991.
3
William K. Rose. Advanced Stellar Astrophysics.
Cambridge University Press, Cambridge (UK) 1998.
44
R. Kippenhahn, A. Weigert. Stellar Structure and Evolution.
Springer-Verlag, Heidelberg und Berlin 1990.
55
S. Chandrasekar. An Introduction the Study of Stellar Structure.
Chicago University Press, Chicago 1931.
66
William E. Boyce, Richard C. DiPrima. Elementary Differential Equations and Boundary Value
Problems. John Wiley and Sons, New York 1997
77
Branden T. Allen. Various Lectures given at Gttingen, Germany on the
polytropic theory of Stars.

4

También podría gustarte