Journal of Modern Physics, 2013, 4, 486-494
http://dx.doi.org/10.4236/jmp.2013.44069 Published Online April 2013 (http://www.scirp.org/journal/jmp)
Scaling Symmetry and Integrable
Spherical Hydrostatics
Sidney Bludman1, Dallas C. Kennedy2
1Departamento de Astronomía, Universidad de Chile, Santiago, Chile
2Natick, USA
Email: sbludman@das.uchile.cl, dalet@stanfordalumni.org
Received October 17, 2012; revised December 10, 2012; accepted December 25, 2012
Copyright © 2013 Sidney Bludman, Dallas C. Kennedy. This is an open access article distributed under the Creative Commons At-
tribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is prop-
erly cited.
ABSTRACT
Any symmetry reduces a second-order differential equation to a first integral: variational symmetries of the action (ex-
emplified by central field dynamics) lead to conservation laws, but symmetries of only the equations of motion (exem-
plified by scale-invariant hydrostatics) yield first-order non-conservation laws between invariants. We obtain these non-
conservation laws by extending Noether’s Theorem to non-variational symmetries and present an innovative variational
formulation of spherical adiabatic hydrostatics. For the scale-invariant case, this novel synthesis of group theory, hydro-
statics, and astrophysics allows us to recover all the known properties of polytropes and define a core radius, inside
which polytropes of index n share a common core mass density structure, and outside of which their envelopes differ.
The Emden solutions (regular solutions of the Lane-Emden equation) are obtained, along with useful approximations.
An appendix discusses the n = 3 polytrope in order to emphasize how the same mechanical structure allows different
thermal structures in relativistic degenerate white dwarfs and zero age main sequence stars.
Keywords: Lagrangian Mechanics; Symmetry; Hydrodynamics; Astrophysics; Stellar Structure
1. Symmetries of Differential Equations and
Reduction of Order
Noether’s Theorem relates every variational symmetry, a
symmetry of an action or similar integral, to a conserva-
tion law, a first integral of the equations of motion [1].
By an extension of Noether’s Theorem, non-variational
symmetries—symmetries of the equations of motion which
are not in general variational symmetries—also lead to
first integrals, which are not conservation laws of the
usual divergence form, as discussed in a previous article
[2]. There it was shown that a Lagrangian
,,tq q

,, dStqq
ii
and action ii , with degrees of freedom
i, can be transformed under an infinitesimal point tran-
sformation :
t
q
 
,, ,
iji
tqq tq




d
d
d
ii i
iii
tq
q q
t
Gqqt
d
dd
,
dd
i
i
qt
q
tt
tt
 


 

 
 







,
i
qqt
(1)
in terms of the total derivative of the Noether charge,
:Gt
ii
p
  and the variational deriva-
tive

:dd
ii i
qqt 
 
 
. For transformations
that leave initial and final states unchanged, the variation
in action is
d
d,
d
f
ifi i
i
SGfGitqt tt



 





(2)
if the term in ddtt
,qt
is integrated by parts. If the
system evolution obeys an action principle, that this va-
riation vanish for independent variations i
0
that
vanish at initial and final times, the system obeys the
Euler-Lagrange equations i and ddtt
 
0
i
,
the rate of change of the Hamiltonian in non-conserva-
tive systems. On-shell, where ,
 
d
f
if i
StGfGi


(3)

d:dd
d
Gtt
t
 
 .
(4)
This is Noethers equation, giving the evolution of a
symmetry generator or Noether charge, in terms of the
Lagrangian transformation that it generates. It expresses
the Euler-Lagrange equations of motion as the diver-
gence of the Noether charge. This divergence vanishes
C
opyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY 487
for a variational symmetry, but not for any other symme-
try transformation.
Noether’s Equation (4) could have been derived direct-
ly from the definition of the Noether charge. But using
the action principle makes manifest the connection be-
tween Noether’s equation and the Euler-Lagrange equa-
tions. We use the action principle and this connection to
reformulate the theory of hydrostatic barotropic spheres,
which is integrable if they are scale symmetric, even
where this scale symmetry is not a symmetry of the ac-
tion (Section 2). The first integrals implied by any sym-
metry of the equations of motion, while generally not va-
nishing-divergence conservation laws, are still useful dy-
namical or structural first-order relationships.
Because it neglects all other structural features, scaling
symmetry is the most general simplification that one can
make for any dynamical system. For the radial scaling
transformations we consider, rr
2;
, the Lagrangian
scales as some scalar density


S
and the ac-
tion scales as . The Noether charge gen-
erating the scale transformation evolves according to a
non-conservation law
12S


12


ddGt , a first-order
equation encapsulating all of the consequences of scaling
symmetry [2]. From this first-order equation follow di-
rectly all the properties of index-n polytropes, as estab-
lished in classical works [3,4], modern textbooks [5,6],
and the recent, excellent treatments of Horedt and Liu
[7,8].
Our secondary purpose is to present an original varia-
tional formulation of spherical hydrostatics and to extend
Noether’s Theorem to non-variational scaling symmetry,
which yields a scaling non-conservation law (Section 2).
For spherical hydrostasis, we define a core radius, inside
which all stars exhibit a common mass density structure.
Outside this core, polytropes of different index n show
different density structures as the outer boundary is felt
(Section 3). Section 4 completes the integration of the
Lane-Emden equation by quadratures and obtains useful
approximations to the Emden function n
.
An appendix reviews the thermodynamic properties of
the physically important polytropes of index n = 3 [2,5,6].
What is original here is the explanation of the the differ-
ences between relativistic degenerate white dwarf stars
and ideal gas stars on the zero-age main sequence (ZAMS),
following from their different entropy structures. Our
original approximations to
3
should prove useful
in such stars.
2. Scaling Symmetry and Integrability of
Hydrostatic Spheres
2.1. Variational Principle for Hydrostatic
Spheres
A non-rotating gaseous sphere in hydrostatic equilibrium
obeys the equations of hydrostatic equilibrium and mass
continuity
2
2
dd ,
dd 4π,
PrGmr
mr r

(5)
where the pressure, mass density, and included mass
Pr,
r
,
mr

depend on radius r. For dependent
variables, we use the gravitational potential
2d
r
VrGmr r

and the thermodynamic potential
(specific enthalpy, ejection energy)

0d
Pr
H
rP
,
so that (5) and its integrated form become
 
dd dd,
,
Hr Vr
GM
Vr HrR
 (6)
expressing the conservation of the specific energy as the
sum of gravitational and internal energies, in a star of
mass M and radius R. The two first-order Equations (5)
are equivalent to a second-order equation of hydrostatic
equilibrium, Poisson’s Law in terms of the enthalpy
H
r:

2
2
1d d4π0,
dd
H
rGH
rr
r
 (7)


We assume a chemically homogeneous spherical
structure, and thermal equilibrium in each mass shell, so
that
r
, Pr,
H
r are even functions of the ra-
dius r. At the origin, spherical symmetry requires
dd 0Pr and mass continuity requires, to order ,
2
r

 
2
33
35
225
1,
4π34π
1.
35 3
c
c
rAr
rr
mr Arr







(8)
The average mass density inside radius r is
 

35
325
:4π3c
rmr rr


0W
.
In a previous paper [2], we showed that hydrostatic
equilibrium (7) follows from the variational principle
minimizing the Gibbs free energy, the integral
of the Lagrangian
:

0
:d,, ,
R
WrrHH

(9)

22
,, 4π8πd,
:dd,
rHHrHG Pr
'r

 
dr
(10)
W is the sum of the gravitational and internal specific
energies per radial shell . The canonical momentum
and
Hamiltonian,
 
2
22 2
:,
,,2 4π,
mHrHG
rHmGmrrP H

 
 
(11)
are the included mass and energy per mass shell. The
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY
488
canonical equations are
2
2
,
4π.
GmrmH
H
mr



 
(12)
Spherical geometry makes the system nonautonomous,
so that 2rrr  vanishes only as-
ymptotically, as the mass shells approach planarity.
The equations of hydrostatic equilibrium (5) can be
rewritten
 
 
dlogdlog =3,
dlogdlog1dlog 1
ururnrvr
vruvr

 dlog ,nr r


(13)
in terms of the logarithmic derivatives

 
 
:dlogd log,
:d logd
: d
urm r
vr P
wr nrvr


log ,
logd log,
r
r

nr
(14)
and an index
 

:dlog
1
1: dlo
nr
nr
d log,
gdlog ,
P
P
 (15)
which depends on the local thermal structure. The mass
density invariant w makes explicit the universal mass
density structure of all stellar cores, which is not appar-
ent in the conventional pressure invariant v.
2.2. Scaling Symmetry and Reduction to
First-Order Equation between Scale
Invariants
Following the our results [2], a hydrostatic structure is
completely integrable, if the structural Equations (5) are
invariant under the infinitesimal scaling transformation


,,
1,
where :21
n
n
rr n
HH
n
,
,
nn
H H
 



 

 


(16)
generated by the Noether charge, for constant n,


2
2
:
4π
2
nn
GrmH
H
rP
 



.
n
H
H
Hr
GG


2
(17)
The Lagrangian (10) then transforms as a scalar den-
sity of weight n

12
12 ,
n
SS
 
n
12
2,
n

 
 (18)
so that only for the polytrope
5

12

n is the
action invariant and scaling a symmetry of the action.
Both structural Equations (13) are autonomous, if and
only if n is constant, so that,

11n
K r
Pr , with
the same constant K (related to the entropy) at each ra-
dius. When this is so1,

ddlog 3,
ddlog 1
n
nnn
uruuw
wrwuwn

 (19)

dlog d logd log
dlog
13
n
nn
wum
r
uwnuwu

 . (20)
In this section, we consider only the first equality in
(20)

1
d
d3
nn
n
n
wu wn
w
uuuw


c
(21)
between scale invariants, which encapsulates all the ef-
fects of scale invariance. We consider only simple poly-
tropes with finite central density
, so that the regular-
ity condition (8) requires that all be tangent to

n
wu

53
3u
at the origin. Such Emden polytropes are the
regular solutions
n
wu

of the first-order Equation (19),
for which

53
3
n
wu u 3 for u.
In terms of the dimensional constant, dimensional ra-
dius, and the central enthalpy and pressure

211
1
1
:,:,
4π
:11 ,,
n
c
n
ccc
c
nKr
G
H
nPnKP



 
(22)
the second-order equation of hydrostatic equilibrium (7),
takes the dimensionless form of the Lane-Emden equa-
tion
22
d
d0.
dd
n
n
n







(23)
In terms of the dimensionless enthalpy nc
H
H

,
the dimensional included mass, mass density, average
included mass density, and specific gravitational force
are

 
 



32
3
2
4π,
,
:3,
4π3
:4π
cn
n
ncn
ncn
cn
mr
r
mr
rr
gr
 







(24)
where prime designates the derivative ':dd
. The
1These characteristic equations are equivalent to a predator/prey equa-
tion in population dynamics [12,13]. With time t replacing-log r, they
are Lotka-Volterra equations, modified by additional spontaneous
growth terms u2, 2
n
wn on the right side. The uw cross-terms lead to
growth of the predator w at the expense of the prey u, so that a popula-
tion that is exclusively prey initially (u = 3, w = 0) is ultimately de-
voured u 0. For the weakest predator/prey interaction (n = 5), the
p
redator takes an infinite time to reach the finite value w5 5. For
stronger predator/prey interaction (n < 5), the predator grows infinitely
in finite time.
n
w
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY
Copyright © 2013 Sci JMP
489
Res.
scale invariants are


11
:,
n
nn
n
n
nn
uv
uv 1
:,
:.
n
n nn
n







(25)
The Noether charge

22
2
21
n
cn
n
H
GGn






1
,
n
nnn



(26)
evolves radially according to

221
d1
21
n
ncnn
n
GH
n
Gn n


 

 
 



 

2
d5
12 .
(27)
This non-conservation law expresses the radial evolu-
tion of energy density per mass shell, from entirely inter-
nal

11
n
nn
at the center, to entirely gravitational
2
2
n

wu
at the stellar surface.
Figure 1 shows the first integrals n for n = 0, 1,
2, 3, 4, 5. For n = 5, scaling is a variational symmetry so
that (26) reduces to a conservation law for the Noether
charge


5
5
1 constant.





v
u
226
255
5
212
3
55
1
262
3
c
c
H
GG
Huvv u
G



  



(28)
For the Emden solution, 5 is finite at the stellar
boundary , the constant vanishes, and
0
 
55
5
53
3
v u 
5nv
wu everywhere.
For , n diverges at the stellar radius 1
, but
0nn
0u
, a finite constant characterizing each Emden
function. At the boundary
, our density invariant
diverges as

u
n
w1
1
0
n
n
n
n
u
, and

3n
2
1

1
01
n
nn
 

0n
. (29)
Table 1 lists these constants
, along with the
global mass density ratios

cn and the ensuing
dimensional radius-mass relation
R


113
14π
n
nnn
M
0n
nKG R

 


 . Together with
the well-known [3,5,6] third, fourth and fifth columns, all
of this table follows directly from the regular solutions of
the first-order Equation (21). In addition, the sixth and
seventh columns express mass concentration in an origin-
nal way.
3. Increasing Polytropic Index and Mass
Concentration
Emden functions are the normalized regular solutions of
the Lane-Emden Equation (23) for which the mass den-
sity is finite at the origin, so that
01, 00


nn
.
Each Emden function of index n is characterized by its
first zero
0

1nn
, at dimensionless boundary radius
1n
. As an alternative measure of core concentration, we
define the core radius core
:2
core
u
implicitly by
,
where gravitational and pressure gradient forces are
maximal. This core radius, where and the mass
density has fallen to
2
n
w
corennc
0.4
1n
for all polytropes
, is marked by red dots in Figures 1-3. The sixth
and seventh columns in Table 1 list dimensionless values
for the fractional core radius corn
rR
ecore 1n
and
fractional included mass coren
m. Within the core
, the internal energy dominates over the gravita-
tional energy, so that for ,
M
2u
1n
 

5
2
core
53,
3
16,for2, ,
n
nn
wu wuu
u
 

 
(30)
consistent with the universal density structure (8) all stars
enjoy near their center.
For n = 0, the mass is uniformly distributed, and the
entire star is core.
As 0 < n < 5 increases, the radial distribution concen-
trates, and the envelope outside the core grows. With
Table 1. Scaling exponents, core parameters, surface parameters, and mass-radius relations for polytropes of increasing mass
concentration. Columns 3 - 5 are well-known [3,5,6]. Columns 6 - 7 present a new measure of core concentration.
Radius-Mass Relation
n n
1n
cn n
R
R

0n
corecore 1nn
rR
coren
mM 31
0
nn
n
RM

0 2 2.449 1 0.333 1 1
R
M; mass uniformly distributed
13
 3.142 3.290 ... 0.66 0.60 1
R
independent of M
1.5 4 3.654 5.991 132.4 0.55 0.51 13
RM
 R
2 2 4.353 11.403 10.50 0.41 0.41
3 1 6.897 54.183 2.018 0.24 0.31 M independent of R
4 2/3 14.972 622.408 0.729 0.13 0.24
4.5 4/7 31.836 6189.47 0.394 0.08 0.22
5 1/2 0 0
0.19
for any
M; mass infinitely
concentrated
S. BLUDMAN, D. C. KENNEDY
490
dlogmudlogr 3
Figure 1. Dilution of polytrope mass density as the boundary is approached 0u. All solutions are tangent to the same
density structure
5n
wz w533u
at the center
3u
, but differ for u < 2 outside the core. Approaching the
outer boundary
0
n
u, the density
nr
falls rapidly, but
1n
n
:
n
n
uv
approaches a constant 1
0
n
n
so that


1
1
0
n
n
nnn
wn u
diverges, for n < 5.
Figure 2. Normalized mass density profiles as a function of fractional included mass mM
, for polytropes of mass concen-
tration increasing with n. The red dots mark the core radii, at which the densities stay near
c
rcore 0.4

0n

n
, for all n 1.
For uniformly distributed mass , the polytrope is all core. As the mass concentration increases , the core
shrinks to about 20% of the mass.
increasing core concentration:
For 1 < n < 3, the radius R decreases with mass M.
Nonrelativistic degenerate stars have 32n
5n

.
For n = 3, the radius R is independent of mass M. This
astrophysically important case is discussed in Section 4
and the Appendix.
For n > 3, the radius R increases with mass M. As
, the stellar radius increases
1315
nnn
 
,
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY 491
Figure 3. Normalized mass density profiles as function of fractional radius r/R. The density is uniform for n = 0, but is maxi-
mally concentrated at finite radius for the n = 5 polytrope, which is unbounded
R
. The density at the core radius stays
about
core c
r0.4
1n, for any .
the core radius shrinks core
103n, the fractional
core radius

0.045 5rR n


corecore 1n,
core 0.19
n
mM, and 01
30
nn


R
.
For n = 5, the mass is infinitely concentrated toward
the center, and the stellar radius for any mass
M
. Scaling becomes a variational symmetry, so that the
Noether charge 5 in (40) is constant with radius. For
the regular solution this constant vanishes:
G


core 5
12
3
55
10 3
31
nG
uvv u
26
55
5 5
1
26 2
0,











(31)
so that 3
55
13, 3vu

 
5. Integrating then yields


12
2
13,

1
5


50
(32)
after normalizing to .
For n > 5, the central density diverges, so that the total
mass M is infinite.
4. Regular Emden Solutions and Their
Approximations
In place of u, we now introduce an equivalent homology
invariant : 3dlogdlogzu r
n
, where

3
:3 4π
nmrr
is the average mass density inside
radius r. In term of z, wn, the characteristic differential
Equations (20) are


dlog
ddlog
g .
3
w
zm
rz

dl
o
32
n
nn
zw z zwn

 (33)
Incorporating the boundary condition, the first of
Equations (40) takes the form of a Volterra integral equa-
tion [9]



 

0
Pic
2
d3
5113: ,
:9 107.
n
zn
nn
n
J
nn
n
zwn
wzzw zw z
J
zwz
Jnn





 
(34)
The Picard approximation is defined by inserting the
core values
53
n
wz z
0, 5n
inside the preceding inte-
gral. For
, this Picard approximation is every-
where exact. For intermediate polytropic indices 0 < n <
5, the Picard approximation breaks down approaching
the boundary, where wn diverges as 1
1
0
n
n
nn
wn u


3n
,
. and is poorest for
After obtaining :dlog dlogr


nn
wz , either nu-
merically or by Picard approximation, another integration
gives [9]

  

0
52
d
exp 3
13
zn
ncn
n
zw z
zwz zz
z







(35)

  

1
0
52
Pic
d
exp 3
13:
n
nn cn
zn
n
n
n
z
zw z
nw zzz
z
 
 





 

(36)

32
3
32
13
exp d
32
3
z
n
z
mz Mzz
wz z
z



 

 
 






(37)
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY
492



1
12
3
12
exp d3
3.
3
n
z
n
rz R
zzzw
z
z







11
32z
z z











RM 3n
(38)
All the scale dependance now appears in the integra-
tion constants M and , which except for
depends on M. Inserting the core values

53wz z
n
inside the integral, the Picard approximations



2
Pic 16,
n
N
nn
NN
 
 :535
n
n 
0, 5n
(39)
to the Emden functions are obtained and tabulated in the
last column of Table 2. For polytropic indices
,
this Picard form is exact. For intermediate polytropic
indices , the Picard approximation remains a
good approximation through order
05n
6
, but breaks down
approaching the outer boundary. Unfortunately, the
Picard approximation is poorest near , the astro-
physically most important polytrope. Figure 4 compares
three approximations to this most important Emden func-
tion, shown in yellow, whose Taylor series expansion is

3n


24 6
3
8
10
164019 5040
619 1088640
2743 39916800.

 

24 6
810
1 0.16666670.0250.0037698
0.00056860.00006872,
(40)
to this Taylor series expansion
 



3core
2.51.7
(41)
shown in red, diverges badly for



.
54
2
3Pic
24 6
12 15
1640 133600
 
 

 
24 6
1 0.16666670.0250.003611,
 
 
3.9
(42)
(43)
shown in dashed green, converges and remains a good
approximation over the bulk of the star, with 10% error
out to
, more than twice the core radius and more
than half-way out to the stellar boundary at 13 6.897
.
nPic
to emden functions
Table 2. Taylor series and picard approximations
.
n
n Emden Function

n
and Taylor Series
:53 5
n
Nn

Picard Approximation

2
Pic :1 6 n
N
nn
N


0 2
16
–1 2
16
1 24 6
sin1 6 1205040
 
  –5/2

52
2246
11516 12010800

 
n
24 6
1612085 15120nnn

 
53 5n


224 6
16161206 510800
n
N
n
Nnnn
 
 
5

12
2
13
1/2

12
2
13
Figure 5. The exact Emden function 3
(solid yellow) and its polynomial (red), Picard (green dashed) and Padé (heavy
black dashed) approximations. Even in this worst case, the Picard approximation holds out to twice the core radius at 2ξ3core =
3.3, before breaking down near the boundary. The Padé approximation is indistinguishable from the exact solution, vanishing
t ξ1 = 6.921, very close to the true boundary at ξ13 = 6.897. a
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY 493
This approximation suffices in white dwarf and ZAMS
stars, except for their outer envelopes, which are never
polytropic and contain little mass. Because it satisfies the
central boundary condition, but not the outer boundary
condition, the Picard approximation underestimates

and overestimates

outside 3.9.
Padé rational approximation [10,11]:
24
3Pad 24
24
1108 1145360
1 171081008
1 0.1666670.025
0.0005686 0.0000857


6
810
0.00376984
618 ,
 

6.921


  

(44)
shown in dashed heavy black, is a simpler and much bet-
ter approximation. By construction, it agrees with the
series expansion (40) through fourth order. In fact, this
Padé approximation is almost exact out to its first zero at
1
13
, very close to the true outer boundary
6.897.
These simple analytic approximations to
3

sim-
plify structural modeling of massive white dwarfs and
ZAMS stars.
5. Conclusions
We have explored how a symmetry of the equations of
motion, but not of the action, reduces a second-order dif-
ferential equation to first-order, which can be integrated
by quadrature. In scale-invariant hydrostatics, the sym-
metry of the equations yields a first integral, which is a
first-order equation between scale invariants, and yields
directly all the familiar properties of polytropes.
We observe that, like all stars, polytropes of index n
share a common core density profile and defined a core
radius outside of which their envelopes differ. The Em-
den functions n
, solutions of the Lane-Emden
equation that are regular at the origin, are finally obtain-
ed, along with useful approximations.
The Appendix reviews the astrophysically most impor-
tant n = 3 polytrope, describing relativistic white dwarf
stars and zero age main sequence stars. While reviewing
these well-known applications [5,6], we stress how these
same mechanical structures differ thermodynamically
and the usefulness of our original (Section IV) approxi-
mations to these Emden functions.
6. Acknowledgements
Thanks to Andrés E. Guzmán (Universidad de Chile) for
calculating the figures with Mathematica and proofread-
ing the manuscript. SAB was supported by the Millen-
nium Center for Supernova Science through grant P06-
045-F funded by Programa Bicentenario de Ciencia y
Tecnología de CONICYT and Programa Iniciativa Cien-
tífica Milenio de MIDEPLAN.
REFERENCES
[1] G. W. Bluman and S. C. Anco, “Symmetry and Integra-
tion Methods for Differential Equations,” Springer-Verlag,
Berlin, 2010.
[2] S. Bludman and D. C. Kennedy, “Invariant Relationships
Deriving from Classical Scaling Transformations,” Jour-
nal of Mathematical Physics, Vol. 52, 2011, Article ID:
042092.
[3] S. Chandrasekhar, “An Introduction to the Study of Stel-
lar Structure, Chapters III, IV,” University of Chicago,
1939.
[4] M. Schwarzschild, “Structure and Evolution of the Stars,”
Princeton University Press, Princeton, 1958.
[5] R. Kippenhahn and A. Weigert, “Stellar Structure And
Evolution,” Springer-Verlag, Berlin, 1990.
[6] C. J. Hansen and S. D. Kawaler, “Stellar Interiors: Physi-
cal Principles, Structure, and Evolution,” Springer-Verlag,
Berlin, 1994.
[7] G. P. Horedt, “Polytropes: Applications in Astrophysics
and Related Fields,” Kluwer, Dordrecht, 2004.
[8] F. K. Liu, “Polytropic Gas Spheres: An Approximate
Analytic Solution of the Lane-Emden Equation,” Monthly
Notices of the Royal Astronomical Society, Vol. 281, No.
4, 1996, pp. 1197-1205.
[9] S. A. Bludman and D. C. Kennedy, “Analytic Models for
the Mechanical Structure of the Solar Core,” The Astro-
physical Journal, Vol. 525, No. 2, 1999, pp. 1024-1031.
[10] P. Pascual, “Lane-Emden Equation and Padé’s Approxi-
mants,” Astronomy & Astrophysics, Vol. 60, 1977, pp.
161-163.
[11] Z. F. Seidov, “Lane-Emden Equation: Picard vs Pade,”
arXiv:astro-ph/0107395.
[12] W. E. Boyce and R. C. DiPrima, “Elementary Differential
Equations and Boundary Value Problems,” 7th Edition,
John Wiley and Sons, Hoboken, 2001.
[13] D. W. Jordon and P. Smith, “Nonlinear Ordinary Differ-
ential Equations,” 3rd Edition, Oxford University Press,
Oxford, 1999.
Copyright © 2013 SciRes. JMP
S. BLUDMAN, D. C. KENNEDY
494
Appendix: Astrophysical Applications of the
n = 3 Polytrope
The polytrope, which is realized in white dwarfs
of maximum mass and in the Eddington standard model
for ZAMS stars just starting hydrogen burning, is distin-
guished by a unique
3n
-
M
R relation: the mass


32
πKG
03
πM
4 is independent of radius R, but
depends on the constant 43
:KP
0WU 
. In these stars, the
gravitational and internal energies cancel, making the
total energy . Because these stars are in
neutral mechanical equilibrium at any radius, they can
expand or contract homologously.
1.1. Relativistic Degenerate Stars: K Fixed by
Fundamental Constants
The most massive white dwarfs are supported by the
degeneracy pressure of relativistic electrons, with num-
ber density eeH
nm

, where
H
m is the atomic
mass unit and the number of electrons per atom
2
eZA
, because these white dwarfs are composed
of pure He or 12 16
CO mixtures. Thus,


13 43
83π
WDH e
Khcm
WD
depends only on funda-
mental constants. This universal value of
K
leads to
the limiting Chandrasekhar mass


22
Ch
2
π815
1.456 2
MM
M
2
5.824
ee
M


[5,6].
1.2. Zero-Age Main Sequence Stars: Mass and
KM Dependent on Specific Radiation
Entropy
In an ideal gas supported by both gas pressure
gas :PTP

 and radiation pressure

3: 1T P

4
rad
Pa , the radiation/gas pressure ratio
is
3
:.
3
Ta
rad
gas
1
P
P


(45)
The specific radiation and ideal monatomic gas entro-
pies are
 

52
l
og,
Tr
r




3
rad gas
4,
3
aT
SSr





(46)
so that the gas entropy gradient

gas
d51
dlog 2ad
ad
S
P


 
 

 
 


(47)
depends on the difference between the adiabatic gradient
25
ad and the star’s actual thermal gradient
:dlog dlogTP

, which depends on the radiation
transport.
Bound in a polytrope of order n, the ideal gas thermal
gradient and gas entropy gradient are

gas
d5
:1 1,1
dlog21
S
nPn

 . 




(48)
For 32n
0.4 150
, the thermal gradient is subadiabatic, the
star’s entropy increases outwards, so that the star is sta-
ble against convection.
M
ZAMS stars, with mass MM

SM
1

, have
nearly constant radiation entropy rad, because
radiation transport leaves the luminosity generated by
interior nuclear burning everywhere proportional to the
local transparency (inverse opacity) . Assuming
constant
SM
3n
rad , we have Eddingtons standard model,
an

polytrope with

rad 41SM



and

13
4
43 31 ,KM Pa



(49)
M
depends only on
, which is itself determined by
Eddingtons quartic equation [3,5,6]
2
2
4
32
03
343
1,
310
:18.3 .
πH
M
M
hc
M
M
Gm








(50)
The luminosity


Edd
3
4
4
Edd
1
0.003 ,
LL M
LMMM





(51)
depends on the Eddington luminosity Edd :4π
p
LcGM
through the photospheric opacity
p
1
. This mass-lumi-
nosity relation is confirmed in ZAMS stars: on the lower-
mass ZAMS,
, ; on the upper-mass
ZAMS,
3
LM

2
21MM

LM, [6].
Copyright © 2013 SciRes. JMP