Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
FREGRAVOL1.pdf
Скачиваний:
71
Добавлен:
05.06.2015
Размер:
5.17 Mб
Скачать

264

 

 

 

 

6

Stellar Equilibrium

and:

 

 

 

 

 

 

 

 

 

4

;

 

 

 

 

kns = 1 ×

 

λn =

 

 

(6.4.28)

 

9π

mnc

As we see (6.4.27) is completely identical in form to (6.4.22). Apart from the minor difference of a factor 2 in the coefficient kns , the quantity Mns is the mass of the neutron star measured in units of baryon masses, just as Mwd was the mass of the white dwarf measured in the same units. Also the factor 2 difference has a nice interpretation. We might write:

k = NB ×

4

(6.4.29)

9π

where, by definition, NB is the number of baryons per fermion of the Fermi gas. In the case of white dwarfs the Fermi gas is made of electrons and for each electron there are two baryons in a star originally made of helium. In the neutron star case, the Fermi gas is made by the neutrons and each of them counts for one baryon. Similarly Rns , just as Rwd , is the radius of the star measured in units of Compton wave-length of the fermionic particles out of which the relevant Fermi gas is made. For neutron stars we measure the radius in neutron wave-lengths just as for white dwarfs we measure it in Compton wave lengths of the electron.

With the above understanding the formula for the Fermi pressure takes a general form valid for both white dwarfs and neutron stars:

 

 

 

 

<

 

 

 

1/3

 

 

 

4

 

 

 

 

1/3

2=

 

 

 

4

M

 

 

 

M

P0 K × λ

×

 

 

 

 

 

 

 

 

 

 

 

 

 

R

 

R

 

 

c

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(6.4.30)

 

 

 

 

 

 

 

 

M

 

 

 

 

 

 

 

 

R

K

;

 

 

M =

 

 

;

 

R =

 

 

 

 

 

 

 

 

12π 2

 

 

 

 

λ

 

 

kmb

 

6.4.2 The Equilibrium Equation

Let us now recall that in Newton’s theory the central pressure of a star of mass M, radius R and constant density is given by (6.3.48). Multiplying and dividing by the same quantities the central pressure can be rewritten in the form given below:

 

3G

×

M2

× λ4 ×

M2

 

Pc =

 

 

= K

 

 

(6.4.31)

8π

R4

R

4

 

 

 

 

 

 

 

 

where by comparison with the original definition (6.3.48), the constant K turns out to have the following value:

K

 

4N 2

Gm2

 

=

B

(6.4.32)

27π 3

 

B

 

6.4 The Chandrasekhar Mass-Limit

 

265

Table 6.1 Table of the values of fundamental constants

 

Name of constant

Symbol

Value

Units

 

 

 

 

Newton constant

G =

6.67428 × 108

cm3 g1 s2

Speed of light

c =

2.99792458 × 1010

cm s1

Planck constant

=

1.054571628 × 1027

g cm2 s2

Proton mass

mp =

1.67262164 × 1024

g

Neutron mass

mn =

1.6749729 × 1024

g

electron mass

me =

0.91093897 × 1027

g

If a neutron-star or a white-dwarf could be conceived as a constant density object, then the equilibrium equation would just be obtained by equating Pc as given in (6.4.31) with the Fermi pressure P0 as given in (6.4.30). Obviously the constant density approximation is too rough, since we know that density varies with depth already in normal stars and even more in compact ones. Yet let us remark that on dimensional grounds, as long as we are in a classical non-relativistic theory, namely as long as we are not supposed to use neither the speed of light c, nor Planck’s constant , the only combination of available parameters G, M and R that has the dimension of a pressure is precisely GMR4 2 , as can be verified by looking at Table 6.1 It follows that solving the pressure differential equation (6.3.37) with any kind of equation of state for the stellar matter:

p = f (ρ)

(6.4.33)

the final result for the central or average pressure in a spherical star of mass M and radius R must be:

P =

3

 

×

M2

 

 

 

8π

R4

 

 

 

 

 

 

 

 

 

 

 

 

 

2

 

 

 

 

 

 

 

 

M

 

 

 

= K ×

λ4

 

4

(6.4.34)

 

 

R

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2N 2

Gαm2

 

K

=

B

 

27π 3

 

 

 

 

 

 

B

 

where α is a dimensionless, numerical coefficient. A very reasonable physical assumption, justified by numerical solutions of the pressure equation in presence of typical equations of state like:

p = γ ; k, γ R

(6.4.35)

is the following: the coefficient α 1 is of order unity. In this case we can draw very important and interesting conclusions. Equating the average gravitational pressure (6.4.34) with the Fermi pressure (6.4.30), we obtain a determination of the star

266 6 Stellar Equilibrium

radius in the form:

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

5

 

 

 

 

 

2/3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1/3

1

M

 

(6.4.36)

 

 

 

 

 

 

 

R

 

M

 

 

 

 

 

 

 

 

 

 

=

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M0

 

 

 

 

 

where we have defined:

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M0

 

K 3/2

 

 

9 π 3/2 c

3/2

1

3/2

 

 

 

=

 

 

 

 

 

=

 

 

 

 

 

 

 

Gm2

 

 

 

 

 

 

 

 

K

 

 

 

8

 

 

 

 

 

 

 

2

α

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

p

 

 

NB

 

 

 

=

1.46386

×

1058

 

 

1

 

3/2

 

 

(6.4.37)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

×

 

 

NB2 α

 

 

 

 

 

The pure number M0 is a mass measured in units of value of such a mass, expressed in grams is:

M0 = M0 4NB mp = 3.87275 × 1033NB

9π

4NB mp . The corresponding

9π

3/2

1

g

NB 2α

=

3.46389

×

NB

1

3/2

×

M

,

(6.4.38)

NB 2α

 

 

 

 

 

where M, = 1.98892 × 1033 g is the mass of the sun.

Formula (6.4.38) holds true for both white dwarfs and neutron stars, the only difference being that in the first case NB = 2, while in the second we have NB = 1. In both cases, as long as α is of order unity also M0 is of order unity in terms of solar masses. What is the meaning of M0? This is evident from (6.4.36). For masses M larger than M0 the radius of the star becomes imaginary, namely the star cannot exist. So M0 is the upper limit for the size of a white-dwarf or a neutron star. It is named the Chandrasekhar mass limit after its discoverer. With minor differences due to α and NB the mass limit for both types of compact stars is approximately the same. Careful considerations on the equation of state of white dwarfs yield for

instance:

 

M0 = 1.4M,

(6.4.39)

which corresponds to a value α 0.46 of the parameter α. We will give a derivation of this result in the next section. The actual radius of white-dwarfs and of neutron stars is however very different. The number R turns out to be approximately the same in both cases, but in the first it measures the radius of the star in units of the Compton-wave length of the electron, while in the second measures it in units of the Compton wave length of the neutron

λe

=

3.861591011

cm

;

λn

=

2.100191014

cm

(6.4.40)

 

 

 

 

 

 

 

Hence the ratio between the white-dwarf radii and those of neutron stars is typically

a factor 103. Since R M1/3 1058/3, for white dwarfs and for neutron-star we respectively have:

6.4 The Chandrasekhar Mass-Limit

 

267

Rwd 108 cm;

Rns 105 cm

(6.4.41)

White-dwarfs are therefore objects with about one solar mass concentrated in the volume of a small planet (thousand kilometer of radius), while neutron-stars have the same mass-magnitude squeezed into a sphere some kilometer-wide.

6.4.3 Polytropes and the Chandrasekhar Mass

Let us now elaborate a little further on the Newtonian equilibrium equation we considered in previous sections. Equation (6.3.37), written for the case of constant density, can be easily generalized to a radial dependent density in the way presented in (6.1.1).

As already stressed this integral-differential equation can be solved if we supply the additional information of an equation of state, relating pressure to energy density. A class of equations of states that generalizes the well known equation of state of ideal gases p = kB ρ is the following one:7

p = κργ ;

1

 

γ = 1 + n

(6.4.42)

Stars whose constituent matter follows an equation of state of type (6.4.42) are named polytropes and the number n is named the polytropic index. In this case the equilibrium equation (6.1.1) can be reduced to a second order non-linear differential equation whose solutions can be determined numerically and in some cases also in closed analytic form. Let us change variables by setting:

= n

ρ(r) ρ0 θ (r) (6.4.43)

where, by definition ρ0 is the energy density at the center of the star and θ (r) is a function of the radial coordinate r whose boundary condition is θ (0) = 1. Similarly, in agreement with (6.4.42), let us set:

= n+1; = 1+ 1

p(r) Pc θ (r) Pc κρ0 n (6.4.44)

where Pc is the central pressure of the star. By means of trivial manipulations the equilibrium equation (6.1.1) becomes:

r

2 d

θ (r) = −

G

 

ρc

M(r)

(6.4.45)

 

 

 

 

 

dr

n +

1 Pc

7In Classical Thermodynamics one writes the relation pV = kB N T where kB is the Boltzmann constant, N the Avogadro number, T the temperature, p the pressure and V the volume. Dividing by V the quantity ρ = NVT can be interpreted as the internal energy density, since the temperature

measures the average kinetic energy per particle and NV is the density of particles.

268

6 Stellar Equilibrium

By taking a second derivative, (6.4.45) is reduced to:

 

1 d

 

 

 

 

 

 

 

 

r2

 

 

= −β

2θ n

(6.4.46)

 

r2

 

dr

dr

where we have defined:

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

β

5

 

(n + 1)Pc

 

 

(6.4.47)

 

 

 

 

4π Gρc2

 

 

 

 

 

 

 

Introducing a rescaled variable z = r/β (6.4.46) is turned into the following standard form:

1 d

z2

+ θ n = 0

(6.4.48)

z2

 

dz

dz

which goes under the name of Lane-Emden equation. The boundary conditions that correspond to the considered physical situation are easily found:

θ (0) = 1;

θ (0) = 0

(6.4.49)

The solution to (6.4.48)–(6.4.49) cannot be written in an explicit analytic form for generic values of n, admitting only a numerical determination. Yet for a few values of n θ (z) is an elementary function. In particular, as the reader can check, we have:

n

θ (z)

 

 

 

 

z0

I0[n]

 

 

 

 

 

 

 

 

 

(

 

)3

 

0

1

1

2

 

6

 

 

 

6 z

 

 

6

3

 

 

1

 

sin(z)

 

 

 

 

(6.4.50)

 

 

 

π

π

 

 

z

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

5

1

 

 

 

 

 

 

 

 

 

 

 

 

1+

z2

 

 

 

 

3

 

 

where we have named z0 the first zero of the θ (z) function and I0 the following integral:

I0

[

n

] ≡

z0

z2

θ (z)

 

n dz

;

θ (z0)

= −

z0θ

(z0)

(6.4.51)

 

 

0

 

 

 

 

 

 

 

For other values of n the above data have to be determined numerically through a computer programme. We also remark that the second identity in (6.4.51) follows from the fact the θ (z) is assumed to satisfy the differential equation (6.4.48).

What is the relevance of the integral I0? This is easily understood, if we calculate the mass M of the star which is supposedly described by the considered equation of state. We immediately find:

M

4π

R r

2ρ(r) dr

=

4πβ3ρ0

z0

z2θ (z) dz

 

 

0

 

 

 

 

 

 

 

0

 

 

=

4π

 

(n + 1

3/2

ρ

3nn

I0

[

n

]

(6.4.52)

 

 

0

 

 

 

4π G

 

 

 

 

 

6.4 The Chandrasekhar Mass-Limit

269

The above relation follows from the simple consideration that the first zero of θ (z), namely z0, defines the radius R βz0 of the star and shows that something very special happens for the polytropic index n = 3. Indeed for such a value the total mass of the star is independent from the central density ρ0. We are therefore led to consider whether n = 3 has a special theoretical interpretation. This is indeed the case. For n = 3 the equation of states takes the form:

p = κρ4/3

(6.4.53)

which is precisely the limiting one for an extremely high density degenerate gas of fermions. How to see it? This is simple. At very high density (xF - 1) we can drop the subleading term in (6.4.14) keeping only the fourth order one. Relying on the form (6.4.11) of the Fermi momentum we conclude that p ρf4/3, where ρf is the density of the fermionic particles composing the gas, electrons in the white dwarf case, neutrons in the neutron-star case. If NB is the number of baryons per Fermi-particle, then we have:

ρ

ρf (6.4.54)

NB mp

and we conclude that p ρ4/3. So we conclude that the equation of state for degenerate compact stars of very high density is indeed of the polytropic type with polytropic index n = 3 γ = 43 . So we can calculate the mass M from formula (6.4.52) if we evaluate the index κ from (6.4.14) and (6.4.11). By direct substitution we find:

 

1

3

 

 

 

1

 

4/3

 

 

 

 

 

 

κ =

 

3

2/3 ×

mp Nb

 

(6.4.55)

4

which inserted in (6.4.52) yields:

 

=

1

 

 

c

3/2

 

1

2I0

3

(6.4.56)

MCh

3π

2

 

NB mp

 

 

 

 

G

 

[ ]

 

which is the celebrated Chandrasekhar formula for the upper mass limit of whitedwarfs and neutron stars. To evaluate it explicitly we just need the value of I0[3] which can be calculated from the numerical solution of the Lane-Emden equation in the case n = 3 (see Fig. 6.11). We find:

z0[3] 6.8;

I0[3] 2.01

(6.4.57)

Inserting this value of I0[3] and those of the fundamental constants in (6.4.56) we find:

MChwd 1.42M,; MChns 5.72M,

(6.4.58)

where we have denoted by MChwd the upper mass-limit for white dwarfs and by MChns the same for neutron stars.

Соседние файлы в предмете [НЕСОРТИРОВАННОЕ]