-
Notifications
You must be signed in to change notification settings - Fork 3
/
232-Notes.tex
5583 lines (5376 loc) · 230 KB
/
232-Notes.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\documentclass[10pt]{article}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% Package Inclusion and Document Formatting %
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\usepackage
{geometry,amsmath,amsthm,mathrsfs,amssymb,graphicx,bm,hyperref,url,pdfsync,
fancyhdr, wrapfig, float}
\pagestyle{fancy}
\numberwithin{equation}{section}
%%%%%%%%%%%%%%%%%%%
% Custom Commands %
%%%%%%%%%%%%%%%%%%%
\newcommand{\n}{\noindent}
\newcommand{\norm}[1]{\left\lvert#1\right\rvert}
\newcommand{\avg}[1]{\left\langle#1\right\rangle}
\newcommand{\abs}[1]{\left\vert#1\right\vert}
\newcommand{\figref}[1]{Figure \ref{#1}}
%%%%%%%%%%%%%%%%%%%%%%%%%%
% Title Page Information %
%%%%%%%%%%%%%%%%%%%%%%%%%%
\title{Notes for PHYS 232: Lars Bildsten's Course on Stellar Structure and
Evolution}
\author{Bill Wolf}
\date{\today}
\begin{document}
\vfill\maketitle\vfill \newpage
\tableofcontents \newpage
\section{Introduction}
\emph{Monday, September 30, 2013}
\subsection{The Magnitude Scale} % (fold)
\label{sub:the_magnitude_scale}
The magnitude is a logarithmic scale detailing the relative brightnesses of
objects. It is based on the star Vega ($\approx 8$ pc away). It is defined such
that its magnitude in the $U$, $B$, and $V$ magnitudes is zero. This
corresponds to a flux definition as
\begin{equation}
\label{mag:1} 3.5\times 10^{-20}\ \mathrm{erg\,cm^{-2}\,s^{-1}\,Hz^{-1}}
\end{equation}
for $M = 0$ (where $M$ is the magnitude). Or, in terms of a simple photon
flux, we have
\begin{equation}
\label{mag:2} N_\lambda = 10^3\ \mathrm{\gamma\,cm^{-2}\,s^{-1}\,\AA}
\end{equation}
at the $V$ band for $V=0$. A useful fact to know is that a magnitude
difference of 5 corresponds to a brightness ratio of 100. The classic filters
have centers and widths given approximately by Table~\ref{tab:mag:1}
\begin{table}[b]
\centering
\begin{tabular}{cccc}
Band & $\lambda_0\ (\AA)$ & FWHM\ (\AA) & Flux ($\mathrm{\#\
cm^{-2}\,s^{-1}\,\AA^{-1}}$)\\
\hline
$U$ & 3650 & 680 & 780 \\
$B$ & 4400 & 980 & 1450 \\
$V$ & 5500 & 890 & 1000 \\
$R$ & 7000 & 2200 & 610 \\
$I$ & 9000 & 2400 & 380
\end{tabular}
\caption{Basic filter band information used in optical astronomy.}
\label{tab:mag:1}
\end{table}
\n The \textbf{Absolute Magnitude} is the apparent magnitude of the star if
it were located at a distance of 10 pc. A \textbf{parsec} (pc) is the distance
at which a star appears to move one arcsecond in the sky over a half of a year
due to the motion of the earth around the sun. In cgs, it's approximatione
\begin{equation}
\label{eq:pc:1} 1\ \mathrm{pc} \approx 3\times 10^{18}\ \mathrm{cm}
\end{equation}
% subsection the_magnitude_scale (end)
\subsection{The HR Diagram}
Most stars shine predominantly in the optical range of the
electromagnetic spectrum. As a result, we get most of our information about
stars by observing their optical output. It makes sense, then, that we
might organize stars by there color, which is indicative of their surface
temperature. When plotting a population of stars' luminosities against
their surface temperature, we note a strong correlation between the two. As
it turns out, the controlling parameter for these quantities is the mass of
the star, at least while the star is on the \textbf{Main Sequence} (stars
burning hydrogen to helium). The correlation between the mass of a main-
sequence star and its luminosity is incredibly strong (see HR diagram
examples).\\
\n Surveys of all nearby stars within 22 pc have found 2241 stars, which
corresponds to a number density of stars as
\begin{equation}
\label{eq:hr:1} n_* = \frac{2241}{\frac{4\pi}{3}(22\ \mathrm{pc})^2} = 0.05\
\mathrm{pc}^{-3}
\end{equation}
which gives an average separation of $\avg{r}\approx 1.7\ \mathrm{pc}$
\begin{figure}[H]
\centering
\includegraphics[width=6in]{figures/hr_local.pdf}
\caption{An HR diagram for stars in the local neighborhood
(shamelessly stolen from Google Images)}
\label{HR.1f}
\end{figure}
\subsection{Conditions for a Star on the HR Diagram}
We are interested in knowing what defines the regime where a star can
reside in a particular $L,\,T_{\mathrm{eff}}=[L/(4\pi\sigma_{\mathrm{SB}}
R^2)]^{1/4}$. Why, for example, is there a dynamical range in the
luminosity spanning over six orders of magnitude, while only a range of a
factor of about 5 in the effective temperature? To gain some perspective,
we might observe the number of stars as a function of brightness. We
organize these stars by their \textbf{spectral type} (a rough measure of
how big the star is) and find their approximate \textbf{mass density} (the
amount of mass contained in these stars per unit volume):
\begin{center}
\begin{tabular}{l l}
Spectral Type & $\rho\ (M_\odot/\mathrm{pc^3})$\\
\hline
O-B & 0.4\\
A-F & 4\\
G-M & 40\\
WD's & 20
\end{tabular}
\end{center}
Here we've used the standard labels for different spectral types, O, B,
A, F, G, K, M, L, T, which are roughly in decreasing order of size and
temperature (the reasoning for this scale is historical rather than
logical, and the ordering is often remembered by the mnemonic, ``Oh Be A
Fine Gal/Guy, Kiss Me! Less Tongue!''). We see that the big, bright stars
form an exceedingly small portion of the amount of stellar mass in our
galaxy. We will find that this is because large stars exhaust their fuel
much more quickly than smaller stars, and thus live and die much faster.
We'll now observe another type of classification of stars used in our
neighborhood, the Milky Way Galaxy.
\subsubsection{Population I Stars}
From Earth, the center of the galaxy is approximately 8.5 kpc away.
Meanwhile, the disk is only about 100 pc wide. We've observed that stars
in the thin disk (commonly known as \textbf{Population I Stars}) are
orbiting at the orbital velocity with only a small amount of axial and
radial motion. They are essentially dynamically cold and in nearly circular
orbits. This is indeed where most of the \textbf{interstellar medium} (ISM)
resides, causing much of the star formation in the galaxy. This region is
also very metal rich. That is, compared to other parts of the universe,
there is a much higher concentration of elements heavier than helium
present. We will denote the mass fraction of these ``metals'' with the
letter $Z$, and in this region, we have $Z\sim 1-2\,\%$. These metals come
from a previous generation of stars, who died in the past, giving off the
metals we now have.
\subsubsection{Population II Stars}
\textbf{Population II Stars} reside mostly in the spheroid in the
center of the galaxy. These are older stars in regions where star formation
is largely shut down. Typically they are metal poor, with metallicities as
low as $Z=10^{-4}Z_\odot$. Kinematically, they are often on radial orbits
(rather than their more azimuthal population I counterparts in the disk).
We typically say that the globular clusters are part of this population.
Sometimes these stars are seen passing through the disk at velocities
comparable to the orbital velocities, and are easily identified due to
their high velocities and unique spectra (due to the low metallicities).
%%%%%
\section{A Simple Hydrostatic Atmosphere}
\subsection{Scale Height and Column Depth}
Before tackling the physics of stars, we first consider a simple toy
model-- the isothermal, plane parallel atmosphere. This model is somewhat
applicable to the thin stellar atmosphere near the surface of a star, where
curvature can be neglected and the acceleration due to gravity is nearly
uniform.\\
\n Consider an atmosphere where the local acceleration due to gravity,
$\mathbf{g}$ is constant in value and direction. The atmosphere is composed
of an isothermal ideal gas with temperature $T$. We wish to find the
distribution of particles in this atmosphere.\\
\n In a strictly statistical sense, we would expect the energy
distribution to be comparable to $e^{-E/kT}$ (recall that the atmosphere is
isothermal, so the average kinetic energy is uniform throughout). In our
case, the energy of particles is linear in height, so we expect this
probability to be proportional to $e^{-mgh/kT}$.\\
\n We will let $m_B\approx m_p$ be the baryon mass, $\mu$ be the mean
molecular mass (measured in AMU), and $\rho$ is the density in $\mathrm{g
\,cm^{-3}}$. We suppose that the gas is in hydrostatic balance, so we have
\begin{equation}
\label{ippa.1} \frac{dP}{dz}=-\rho g
\end{equation}
Combining this with the ideal gas law,
\begin{equation}
\label{ippa.2} P=nkT
\end{equation}
We find that
\begin{equation}
\label{ippa.3} kT\frac{dn}{dz}=-m_p\mu ng
\end{equation}
which in turn gives us
\begin{equation}
\label{ippa.4} \frac{d\ln n}{dz}=-\frac{m_p\mu g}{kT}
\end{equation}
Solving this differential equation gives the expected result
\begin{equation}
\label{ippa.5} n(z)=n(0)\exp\left(-\frac{m_p\mu gz}{kT}\right)=n
(0)\exp\left(-\frac{z}{h}\right)
\end{equation}
where we have defined the \textbf{scale height} $h\equiv kT/(\mu m_pg)
$, which is the $e$-folding distance in number density. As it turns out, the
scale height for earth's atmosphere is approximately 10 km. This model is
really only valid in cases where $h\ll R$, (where $R$ is the size of the
object), so we now investigate the ratio of these two quantities:
\begin{equation}
\label{ippa.6} \frac{h}{R}=\frac{kT}{\mu m_p\frac{GM}{R^2}R}=\frac
{1}{\mu}\frac{kT/m_p}{GM/R}\sim \frac{v_{\mathrm{th}}^2}{v_{\mathrm{esc}}
^2}\ll 1
\end{equation}
So if this approximation is valid, the thermal velocities of particles
are typically much smaller than the escape velocity of the central body, so
a star could retain its own atmosphere (thankfully, Earth does the same to
its atmosphere!) For stars, we will find that $kT_c/m_p\sim GM/R$. Then,
\eqref{ippa.6} tells us that
\begin{equation}
\label{ippa.7} \frac{h}{R}\sim \frac{T_{\mathrm{eff}}}{T_c}.
\end{equation}
This tells us that stars are quite sharp-edged (their scale heights are
very small compared to their radii). We can also deduce a physical meaning
for the scale height as being how far a particle needs to fall to gain an
energy comparable to $kT$.\\
\n From the ideal gas law, we can easily see that the pressure will
also fall off exponentially in this isothermal atmosphere. However, let's
explore the pressure a bit more. First, we return to the ideal gas law,
\begin{equation}
\label{ippa.8} P=\frac{\rho}{\mu m_p}kT=nkT
\end{equation}
and the condition for hydrostatic equilibrium,
\begin{equation}
\label{ippa.9} dP=-\rho g\,dz.
\end{equation}
We now integrate \eqref{ippa.9} from $z=z$ to $z\to \infty$:
\begin{align}
\label{ippa.10} P(\infty)-P(z)&=-\int_z^\infty \rho(z') g\,dz'\\
\label{ippa.11} P(z) &= g\int_z^\infty \rho(z')\,dz
\end{align}
Note that it is okay to take the integral to infinity so long as we are
dealing with a constant $\mathbf{g}$. This result suggests the definition
of the \textbf{column density}:
\begin{equation}
\label{ippa.12} y(z)\equiv \int_z^\infty \rho(z)\,dz
\end{equation}
On the surface of the earth, the column density is approximately
$y=1000\,\mathrm{g\,cm^{-2}}$. Think of it as the amount of mass sitting
above you per unit area at a given altitude. The column density is an
important number (for us) to determine the details of heat transport. For
now though, we can write the pressure in this isothermal atmosphere in a
compact form: $P(z)=gy(z)$.
\subsection{Mean Molecular Weights}
We'll now make a useful definition for calculating pressures and other
useful quantities. For an ideal gas, the total pressure of a mixed gas is
simply
\begin{equation}
\label{mmw.1} P=\sum_{i=1}^N n_ikT
\end{equation}
where $n_i$ are just the number densities for each ion. The number
density is computed via
\begin{equation}
\label{mmw.2} n_i=\frac{X_i\rho}{A_im_p}.
\end{equation}
where $X_i$ is the mass fraction of the $i^{\mathrm{th}}$ ion with mass
number $A_i$ and $\rho$ is the overall mass density. Then the ion pressure
is given by (assuming total ionization)
\begin{equation}
\label{mmw.3} P_{\mathrm{ion}}=kT\sum\frac{X_i\rho}{A_im_p}=\frac
{kT\rho}{m_p}\sum\frac{X_i}{A_i}=\frac{kT\rho}{\mu_{\mathrm{ion}}m_p}.
\end{equation}
Where we have defined the \textbf{mean molecular weight} of the ions to
be
\begin{equation}
\label{mmw.3a} \mu_{\mathrm{ion}}=\left[\sum\frac{X_i}{A_i}\right]^
{-1}
\end{equation}
For the electrons, we have (assuming total ionization)
\begin{equation}
\label{mmw.4} P_e=n_ekT=kT\left(\sum Z_in_i\right)=\frac{kT\rho}
{m_p}\sum\frac{Z_iX_i}{A_i}.
\end{equation}
(here $Z_i$ is the atomic number, not the metallicity). Then the total
pressure is just the sum of these two,
\begin{equation}
\label{mmw.5} P=P_{\mathrm{ion}}+P_e=\frac{\rho kT}{m_p}\left(\frac
{1}{\mu_e}+\frac{1}{\mu_i}\right)
\end{equation}
So we define the overall mean molecular weight via
\begin{equation}
\frac{1}{\mu}\equiv \frac{1}{\mu_e}+\frac{1}{\mu_i}
\end{equation}
For fully-ionized solar composition material, we have $\mu \approx 0.64$.\\
\n One might think of this as the average weight of a particle that
supplies pressure within a gas. Later, we'll see that this quantity, and
its evolution, plays a large and critical role in the the nature of stellar
evolution. Since fusion tends to decrease the pressure support, the star
must continuously readjust its structure so as to hold itself up.\\
%%%%%%%%%%%%%%%%%%%%%%
% October 2, 2013 %
%%%%%%%%%%%%%%%%%%%%%%
\n \textit{Wednesday, October 2, 2013}
\subsection{Electric Fields in Stars}
Imagine a pure, ionized hydrogen atmosphere which is, on the large
scale, electrically neutral. We wish to find the scale height in a plasma
of ionized hydrogen. In this plasma, we have $n_p=n_e$ due to electric
neutrality. Then the overall pressure in this hydrogen plasma is
\begin{equation}
\label{efs.2} P=2n_pkT
\end{equation}
Using hydrostatic equilibrium, we get
\begin{equation}
\label{efs.3} 2kT\frac{dn_p}{dz}=-m_pn_pg
\end{equation}
which in turn gives us the differential equation
\begin{equation}
\label{efs.4} \frac{d\ln n_p}{dz}=-\frac{m_pg}{2kT}
\end{equation}
Which gives us a scale height of
\begin{equation}
\label{efs.5} h=\frac{2kT}{m_pg}
\end{equation}
We need to look at both plasmas separately while incorporating the
electric field created between the protons and electrons. For electrons, we
have
\begin{equation}
\label{efs.6} \frac{1}{n_e}\frac{dP_e}{dz}=-m_eg-eE.
\end{equation}
Likewise for the protons,
\begin{equation}
\label{efs.7} \frac{1}{n_p}\frac{dP_p}{dz}=-m_pg+eE
\end{equation}
where we've assumed that the electric field points up (the protons are
heavier and would thus sink below the electrons). Now adding \eqref{efs.6}
and \eqref{efs.7}, we recover hydrostatic balance. However, subtracting the
two equations will get us the electric field:
\begin{equation}
\label{efs.8} 0=-m_eg+m_pg-2eE,
\end{equation}
giving the result,
\begin{equation}
\label{efs.9} eE=\frac{1}{2}\left(m_p-m_e\right)g \quad \textrm{or}
\quad e\mathbf{E}\approx -\frac{m_p\mathbf{g}}{2}.
\end{equation}
So this field does not directly affect hydrostatic balance, but it does
dramatically impact the relative surface abundances of elements since different isotopes feel different forces, causing diffusion and sedimentation.
\section{Self-Gravitating Objects}
So far we have only considered systems where the acceleration due to
gravity is constant. In any self-gravitating object, this is obviously not
true. We will, however, continue to assume that such objects do not rotate.
Additionally, we will be ignoring mass loss. Essentially all we must write
down are equations of mass conservation, momentum conservation, and energy
conservation. We'll start with momentum conservation.
\subsection{Momentum Conservation and the Free-Fall Timescale}
The momentum equation for a fluid is just
\begin{equation}
\label{mc.1} \rho\frac{d\mathbf{v}}{dt}=\rho\mathbf{g}-\bm{\nabla}P
\end{equation}
This equation essentially states that a self-gravitating object is
neither collapsing nor expanding. If we were to ``shut off'' gravity or the
pressure gradient, the star would either explode or collapse, respectively.
Such a collapse would occur on the \textbf{free-fall timescale}, which we
will now derive. Taking the pressure gradient out of \eqref{mc.1}, we
retrieve
\begin{equation}
\label{mc.2} \mathbf{g}=-\frac{Gm(r)}{r^2}\hat{r}
\end{equation}
For this derivation, we will be using a \textbf{Lagrangian coordinate
system}. This is a system where the coordinates follow a particular fluid
element. In essence, we are making the substitution
\begin{equation}
\label{mc.3} \frac{d}{dt}\to\frac{\partial}{\partial t}+\mathbf{v}
\cdot\bm{\nabla}
\end{equation}
Returning back to the derivation, \eqref{mc.2} gives us
\begin{equation}
\label{mc.4} \frac{dv_r}{dt}=-\frac{Gm(r)}{r^2}
\end{equation}
Initially, we have $t=0$, $v_r=0$, and $r=r_0$ with the radial velocity
given by $v_r=dr/dt$. Then our differential equation is
\begin{equation}
\label{mc.5} \frac{d^2r}{dt^2}=-\frac{Gm(r)}{r^2}
\end{equation}
As an order of magnitude estimate, this gives us
\begin{equation}
\label{mc.6}\frac{r}{t_{\mathrm{ff}}^2}\sim\frac{Gm}{r^2}\quad
\Rightarrow \quad t_{\mathrm{ff}}^2\sim\frac{1}{Gm/r^3}
\end{equation}
So we define the free-fall timescale to be
\begin{equation}
\label{mc.7} t_{\mathrm{ff}}=\frac{1}{\sqrt{G\avg{\rho}}}
\end{equation}
This is also the same as the Keplerian orbital period, modulo some
uninteresting constants. The punchline of this whole argument is that a
star that is \emph{not} in hydrostatic balance will respond on a timescale
of the free-fall timescale. From this alone, we may conclude that the sun
(and any other star not currently exploding) is in hydrostatic balance. We
will then assume that all stars are always in hydrostatic balance.\\
\subsection{The Virial Theorem}
Stars are held up by gas pressure, radiation pressure, or both. The
pressure gradients are what will be the ``restoring forces'' against
gravity for our cases. In spherical symmetry, hydrostatic balance tells us
\begin{equation}
\label{rss.1} \frac{dP}{dr}=-\rho\frac{Gm(r)}{r^2}
\end{equation}
We will use this to derive the \textbf{Virial Theorem}, which relates
the potential energy to the kinetic energy of a system. The equation of
mass conservation states that
\begin{equation}
\label{rss.2} dm=4\pi r^2\rho(r)\,dr
\end{equation}
Now we multiply both sides of \eqref{rss.1} by $4\pi r^3\,dr$:
\begin{align}
\label{rss.3} \int 4\pi r^3\,dP&=-\int \rho\frac{Gm(r)}{r^2}4\pi
r^3\,dr\\
\label{rss.4} &= -\int\frac{Gm(r)dm}{r} = E_{\mathrm{GR}}
\end{align}
where $E_{\mathrm{GR}}$ is the gravitational binding energy. Performing
a similar analysis to the left side of \eqref{rss.3} gives
\begin{align}
\label{rss.5} \int 4\pi r^3\,dr\frac{dP}{dr} &= \left.4\pi r^3P
\right|_{0}^R-3\left[4\pi\int Pr^2\,dr\right]\\
\label{rss.6} &= -3\int P\,4\pi r^2\,dr\\
\label{rss.7}&=-3\avg{P}V
\end{align}
where we've defined the average pressure to be the pressure
averaged over volume. Also, in \eqref{rss.6} we've noted that
the boundary terms must vanish since $P(R)=0$ (not necessarily the case
during star formation in dark clouds, though). Then the virial
theorem tells us that
\begin{equation}
\label{rss.8}\boxed{\avg{P}=-\frac{1}{3}\frac{E_{\mathrm{GR}}}{V}}
\end{equation}
Now we examine the total energy:
\begin{equation}
\label{rss.9} E_{\mathrm{tot}}=E_{\mathrm{GR}}+E_{\mathrm{KE}}
=-3\avg{P}V+E_{\mathrm{KE}}
\end{equation}
We need only relate the kinetic energy to the pressure to finish this
equation off. For an ideal gas, we know that $P=NkT/V$, so the kinetic
energy is $E_{\mathrm{KE}}=\frac{3}{2}NkT=\frac{3}{2}PV$. This gives a
total energy of
\begin{equation}
\label{rss.10} E_{\mathrm{tot}}=-3\avg{P}V+\frac{3}{2}\avg{P}V=-E_
{\mathrm{KE}}
\end{equation}
Interestingly, this requires a negative heat capacity. That is, an
increase in the temperature of the system causes a net \emph{decrease} in
total energy. However for radiation, pressure is given by $P=\frac{1}{3}
aT^4$ and $E/V=aT^4$. Taking this to its conclusion gives us
\begin{equation}
\label{rss.11} E_{\mathrm{tot}}\to 0\ \textrm{as the particles
become relativistic}
\end{equation}
The origin of this result is in the momentum-energy relation of
relativistic particles and non-relativistic particles. That is, $E=pc$ for
ultra-relativistic particles and $E=p^2/2m$ for non-relativistic particles.
\\
\n The limiting energy of ultra-relativistic stars puts an upper level
on the mass of large stars, since a total energy of a star being zero means
unbinding the star. In the ``normal case'' of an ideal gas star, the more
traditional form of the virial theorem applies:
\begin{equation}
\label{rss.12} \frac{E_{\mathrm{KE}}}{\mathrm{mass}}\sim\frac{GM}
{R}
\end{equation}
This is why stars typically behave with a negative heat capacity. That
is, as a star radiates, $E_{\mathrm{tot}}$ is more negative, meaning that
$R$ must decrease and the temperature $T$ (essentially the kinetic energy
per particle) rises. This behavior would have to continue until a new
energy source became available.
\subsection{Applications of the Virial Theorm}
The gravitational energy of an object is typically given by
\begin{equation}
\label{avt.1}E_{\mathrm{GR}}\approx -\frac{GM^2}{R}
\end{equation}
Using the virial theorem, we have
\begin{equation}
\label{avt.2} -E_{\mathrm{GR}}=3\avg{P}V=3Nk\avg{T}
\end{equation}
Or,
\begin{equation}
\label{avt.3} \frac{GM}{R}\left(Nm_p\right)\sim 3NkT
\end{equation}
So we have
\begin{equation}
\label{avt.4} \boxed{kT\sim \frac{GMm_p}{R}}
\end{equation}
This temperature is the temperature of most of the material
and is $T\sim T_c$ (temperature in the core). For the sun, we
then have $T\sim 10^7\ \mathrm{K}$. Interestingly, assuming
hydrostatic equilibrium was all we needed to get a rough
estimate of the sun's core temperature! One might note,
though, that the surface temperature is significantly lower
than the core temperature, so we must assume that there is
heat loss taking place in the sun. Today the luminosity of the
sun is
\begin{equation}
\label{avt.5} L_\odot = 4\times 10^{33}\ \mathrm{erg/s}
\end{equation}
If we assume there is no energy source for the sun other than
gravitational energy, we can come up with a timescale (called the \textbf
{Kelvin-Helmholtz timescale})
\begin{equation}
\label{avt.6} t_{\mathrm{KH}}=\norm{\frac{E_{\mathrm{GR}}}{L}}
\approx 3\times 10^7\ \mathrm{years}
\end{equation}
for the sun. This has been known for awhile and since the Earth is
known to have existed much longer than $t_{\mathrm{KH}}$, scientists
deduced that another energy source within the sun was needed to explain its
longevity. We now know that this energy source is, of course, fusion. Note
that at the center of the sun, the temperature of $10^7\ \mathrm{K}$
corresponds to an energy per particle of about 1 keV. The binding energy of
helium is approximately 7MeV, approximately 7000 times bigger than the
thermal content. Thus, the sun could last approximately 7000 times longer,
bringing the projected lifetime of the sun up to a more reasonable (but
still wrong) number of about 200 billion years. We conclude that nuclear
energy is a more promising form of energy for the sun than chemical energy.
\\
\n\textit{Monday, October 7, 2013 (No updates)}
\subsection{Gas Pressure and Radiation Pressure}
Recall from the case of the constant density star that the
gravitational energy is given by
\begin{equation}
\label{GPRP.1} E_{\mathrm{GR}}=-\frac{3}{5}\frac{GM^2}{R}
\end{equation}
And the average pressure is given by the virial theorem to be
\begin{equation}
\label{GPRP.2} \avg{P}=-\frac{1}{3}\frac{E_{\mathrm{GR}}}{V}=
(n_e+n_p)kT=2n_pkT=\frac{2\rho kT}{m_p}
\end{equation}
Here we've sort of assumed that the star is
isothermal. This tells us that the average thermal
energy is given by
\begin{equation}
\label{GPRP.3} kT=\frac{1}{10}\frac{GMm_p}{R}
\end{equation}
Where the mass is given by
\begin{equation}
\label{GPRP.4} M=\rho\frac{4\pi}{3}R^3
\end{equation}
and the central temperature is given approximately by (scaled by
solar units)
\begin{equation}
\label{GPRP.5} T_c\approx2\times 10^6\,\mathrm{K}\left(\frac
{\rho_c}{1\,\mathrm{g\,cm^{-3}}}\right)^{1/3}\left(\frac{M}{M_\odot}\right)
^{2/3}
\end{equation}
These scalings are actually recovered in simulations
(see Figure~\ref{fig:paxton_2011_contraction}).
\begin{figure}[H]
\includegraphics[width=0.8\textwidth]{figures/paxton_2011_contraction.png}
\caption{From Paxton et al. 2011 (MESA I)}
\label{fig:paxton_2011_contraction}
\end{figure}
Here we've only considered
the case of the pressure due to an ideal gas, thus far
ignoring the contributions from radiation pressure. We
then want to know when radiation pressure becomes
comparable to gas pressure. That is,
\begin{equation}
\label{GPRP.6} P_{\mathrm{rad}}=\frac{1}{3} aT^4\geq P_{\mathrm
{gas}}
\end{equation}
The temperature in the star is approximately
\begin{equation}
\label{GPRP.7} kT\sim\frac{GMm_p}{R}
\end{equation}
and the pressure gradient is, (again, very approximately)
\begin{equation}
\label{GPRP.8} \frac{dP}{dr}=-\rho g\approx \frac{P_c}{R}\sim
\frac{M}{R^3}\frac{GM}{R^2}
\end{equation}
Then the condition we are seeking is
\begin{equation}
\label{GPRP.9} \frac{1}{3}a\left(\frac{GMm_p}{Rk}\right)
^4\gtrsim\frac{GM^2}{R^4}
\end{equation}
Interestingly, $R$ cancels in \eqref{GPRP.9}, so this condition is
dependent only on the mass of the star. Thus, we can get a hard limit that
is independent of any other properties of the star. Dropping tons more
constants, this gives
\begin{equation}
\label{GPRP.10} M^2>\frac{k^4}{a G^3m_p^4}
\end{equation}
Recall that the radiation constant is given by $a=\frac{\pi^2}
{15}\frac{k^4}{(\hbar c)^3}$. Putting this in \eqref{GPRP.10}, we have
\begin{equation}
\label{GPRP.11} \frac{M^2}{m_p^2}>\frac{k^4(\hbar c)^3}
{G^3m_p^6 k^4}\sim\left(\frac{\hbar c}{G m_p^2}\right)^3
\end{equation}
Then the limit on the mass is then
\begin{equation}
\label{GPRP.12}\boxed{ M>m_p\left(\frac{\hbar c}{Gm_p^2}\right)
^{3/2}}
\end{equation}
stars above this mass (approximately) have significant radiation
pressure. Recall the fine structure constant
\begin{equation}
\label{GPRP.13} \alpha=\frac{1}{137} =\frac{e^2}{\hbar c}
\end{equation}
Noting that the Coulomb energy is
\begin{equation}
\label{GPRP.14} E_{\mathrm{Coulomb}}=\frac{e^2}{r}
\end{equation}
Analagous to \eqref{GPRP.14}, we define a
dimensionless measure of the strength of gravity,
which appears in \eqref{GPRP.12}:
\begin{equation}
\label{GPRP.15} \alpha_\mathrm{G}=\frac{Gm_p^2}{\hbar c}\approx
6\times 10^{-39}
\end{equation}
Then the fundamental stellar mass given in \eqref{GPRP.12} is
\begin{equation}
\label{GPRP.16} M>m_p\frac{1}{\alpha_{\mathrm{G}}^{3/2}}\approx
2M_\odot
\end{equation}
After all that work\ldots it turns out that the mass where
radiation pressure \emph{actually} starts to matter is closer to $60-90\ M_
\odot$.\\
\n Of interest in this case is that as $P_{\mathrm{rad}}\gg P_
{\mathrm{gas}}$, then $E_{\mathrm{tot}}\to 0$ from the virial theorem. In
this state, the star has nearly enough energy to unbind itself, so radiation
pressure sets an upper limit on the mass of a star as other instabilities will
act to unbind it when the pressure is radiation dominated.
\subsection{Summary}
Note that here we have used hydrostatic balance to
find the central temperature as a function of mass and
radius. Additionally we have realized that energy
losses from the surface require the radius of a star
to decrease and the core temperature to increase (at
least until another energy source is present). We will
later show that the main sequence is just that place
where the power generated by nuclear reactions is
equal to that released by the star so that the star
need not contract. What we have \emph{not} done yet is to
derive the rate of heat loss from the star.
\section{Heat Transfer in Stars}
In studying the ways in which heat moves outward through a star, we
will first be ignoring convection, though it is a powerful mechanism when
it is available to the star. To move heat through a star, there are
electrons, ions, and photons available. Recall (or perhaps you don't)
\textbf{Fick's Law}, which tells us that the heat flux can be determined
via
\begin{equation}
\label{HTS.1} F=\mathrm{ergs\,cm^{-2}\,s^{-1}}=-\frac{1}{3}v\ell
\frac{d}{dx}(E)
\end{equation}
where here $E$ is the energy density. We'll start first with
the energy density of an ideal gas of electrons, but before we
do that, let's derive \eqref{HTS.1}.
\subsection{Heat Flux Derivation (not done in class)}
Imagine a medium with a gradient in temperature across a surface
membrane. Let's say ``above'' the membrane, the temperature is $T_1$
and ``below'' the membrane, the temperature is $T_2$, with
$T_2>T_1$. Particles from region 2 then transport heat when they
travel from region 2 to region 1. Let's call $E$ the internal energy
per unit volume as defined before, and then let's see what happens at
the surface. Particles, on average, will be coming from a distance
$x+\ell$ above the membrane, where $\ell=(\sigma n)^{-1}$ is the mean
free path of the particles. Then the downward flow of energy is
\begin{equation}
\label{eq:3}
F_{\mathrm{down}}\approx \frac{1}{6} vE(x+\ell)
\end{equation}
whereas particles from beneath move upward and carry heat from below
at
\begin{equation}
\label{eq:1}
F_{\mathrm{up}}\approx \frac{1}{6}vE(x-\ell)
\end{equation}
Think of the factor of $1/6$ as being the portion of the flux moving
through a particular face of a cube, in this case, the face pointing
up or down. So the net flux in the positive $\hat{x}$ direction is
\begin{equation}
\label{eq:1a}
F_x=-\frac{1}{6}vE(x+\ell)+\frac{1}{6}vE(x-\ell)
\end{equation}
Writing the energy densities as linear functions,
\begin{align}
\label{eq:2}
E(x+\ell) &= E(x)+\ell\frac{dE}{dx}\\
\label{eq:2a}
E(x-\ell) &= E(x)-\ell\frac{dE}{dx}
\end{align}
so
\begin{equation}
\label{eq:4}
F_x=-\frac{1}{3}v\ell\frac{dE}{dx}
\end{equation}
\subsection{Heat Transport by Electrons} For electrons, the energy
density is
\begin{equation}
\label{HTS.2} E=\frac{3}{2} kT n_e
\end{equation}
Then the energy density gradient is
\begin{equation}
\label{HTS.3} \frac{dE}{dx}=\frac{dE}{dT}\frac{dT}{dx}=\frac{3}{2}
nk\frac{dT}{dx}
\end{equation}
Then from Fick's Law, we have
\begin{equation}
\label{HTS.4} F=-\frac{1}{3}v\ell\frac{3}{2}nk\frac{dT}{dx}=-\frac
{1}{2}v\ell nk\frac{dT}{dx}
\end{equation}
Where the mean free path is
\begin{equation}
\label{HTS.5} \ell=\frac{1}{n\sigma}
\end{equation}
for the scattering cross section $\sigma$. Then \eqref{HTS.4} becomes
\begin{equation}
\label{HTS.6} F=-\left[\frac{1}{2} v\frac{k}{\sigma}\right]\bm
{\nabla}T
\end{equation}
From the theory of Coulomb scattering, the cross section would be
\begin{equation}
\label{HTS.7} \sigma_{\mathrm{Coulomb}}\sim b^2\sim\frac{e^4}{(kT)
^2}
\end{equation}
Which gives a flux of
\begin{equation}
\label{HTS.8} L=4\pi R^2F=4\pi R\frac{(kT)^{7/2}}{m_e^{1/2}e^4}
\end{equation}
For the sun, this would give
\begin{equation}
\label{HTS.9} L\approx 5\times 10^{31}\,\mathrm{erg\,s^{-1}}
\end{equation}
which is two orders of magnitude too small, so we conclude that the sun
does not transmit its heat via conduction through electrons. Now we'll move
on to photons
\subsection{Radiative Diffusion}
For photons, the main scatterer will be electrons, so the cross section
in question of the mean free path is the Thomson cross section.
Additionally, the energy density is now $E=aT^4$. Additionally the speed of
photons is obviously the speed of light. Then the ratio of the fluxes due
to photons and electrons is
\begin{equation}
\label{HTS.10} \frac{F_{\gamma}}{F_e}=\frac{c}{(kT/m_e)^{1/2}}\frac
{e^4/(kT)^2}{\sigma_{\mathrm{Th}}}\frac{E_{\mathrm{rad}}}{E_{\mathrm{gas}}}
\end{equation}
Remember that the Thomson cross section is given by
\begin{equation}
\label{HTS.11} \sigma_{\mathrm{Th}}=\frac{8\pi}{3}\frac{e^4}
{(m_ec^2)^2}
\end{equation}
Then \eqref{HTS.10} becomes
\begin{equation}
\label{HTS.12} \frac{F_\gamma}{F_e}=\left(\frac{m_ec^2}{kT}\right)^
{1/2}\left(\frac{m_ec^2}{kT}\right)^2\frac{P_{\mathrm{rad}}}{P_{\mathrm
{gas}}}
\end{equation}
Comparing the pressures gives
\begin{equation}
\label{HTS.13} \frac{P_{\mathrm{rad}}}{P_{\mathrm{gas}}}\approx 10^
{-4}\left(\frac{M}{M_\odot}\right)^2
\end{equation}
(this is essentially a more accurate estimate of what we determined in
\eqref{GPRP.16}.) Then plugging \eqref{HTS.13} into \eqref{HTS.12}, we see
that heat transport by photons dominates heat transport by electrons whenever
\begin{equation}
\label{HTS.14} M>0.03\,M_\odot\,\left(\frac{T}{10^7\ \mathrm{K}}
\right)^{5/4}
\end{equation}
So if conduction ever dominates, it is in very low mass stars (also
white dwarfs in their late lives). For our cases, photons are always going
to be the dominant transport mechanism. We still haven't found out what the
actual luminosity will be in the case of radiative diffusion. We do so now.
\begin{equation}
\label{HTS.15} L=4\pi R^2 F=4\pi R^2\frac{1}{3}c\ell\frac{d}{dr}
\left(aT^4\right)\approx R^2c\frac{1}{n_e\sigma_{\mathrm{Th}}}\frac{1}{R}
aT^4\approx R^2\frac{c m_p}{\rho\sigma_{\mathrm{Th}}}\frac{1}{R}a\left
(\frac{GMm_p}{Rk}\right)^4
\end{equation}
where we have noted that $kT=\frac{GMm_p}{R}$. Continuing on,
\begin{equation}
\label{HTS.16} L\approx \frac{cm_pa(GMm_p)^4}{M\sigma_{\mathrm{Th}}
k^4}\propto M^3
\end{equation}
This relation is surprisingly accurate for stars with masses greater
than a solar mass. Note that we have derived the stellar luminosity with
\emph{no knowledge} of the source of energy. The luminosity is set by the
modes of heat transport available to the star.\\
%%%%%
\n \textit{Wednesday, October 9, 2013}
\subsubsection{The Eddington Limit and the Eddington Standard Model}
Continuing with heat transfer via radiation with electron scattering
being the primary source of opacity, the flux is given via Fick's law as
\begin{equation}
\label{HTS.17} F=\frac{1}{3} v\ell\frac{d}{dz}(aT^4)
\end{equation}
where now $v=c$ and $\ell=(n_e\sigma_{\mathrm{Th}})^{-1}$. Plugging
these in to \eqref{HTS.17}, the flux becomes
\begin{equation}
\label{HTS.18} F=\frac{1}{3}c\frac{1}{n_e\sigma_{\mathrm{Th}}}\frac
{d}{dz}(aT^4)=\frac{4}{3}\frac{acT^3}{\rho\kappa}\frac{dT}{dz}
\end{equation}
where we've defined the \textbf{opacity} by generalizing the relationship
between the density and mean free path via
\begin{equation}
\label{HTS.19} \ell = \frac{1}{n \sigma} \equiv \frac{1}{\rho \kappa}
\qquad \Rightarrow \qquad \kappa_{\mathrm{es}}\equiv \frac{n_e \sigma_{\rm Th}}{\rho}
= \frac{\sigma_{\mathrm{Th}}}{\mu_e m_p} \approx 0.4\ \mathrm{cm^2\ g^{-1}}
\end{equation}
The opacity is measured in units of area per unit mass, indicating it
is the cross-sectional area per unit mass. For electron scattering, it is
essentially constant, but when other processes are relevant, it may depend on
temperature and density. With the flux available, we can write the
luminosity as
\begin{equation}
\label{HTS.20} L(r)=F(r)4\pi r^2=-4\pi r^2\frac{4}{3}\frac{acT^3}
{\rho\kappa}\frac{dT}{dr}=-4\pi r^2\frac{c}{\kappa\rho}\frac{d}{dr}P_
{\mathrm{rad}}
\end{equation}
Noting the product $\rho\,dr$ showing up in the denominator of \eqref
{HTS.20}, we are reminded of hydrostatic equilibrium:
\begin{equation}
\label{HTS.21} \frac{dP}{dr}=-\rho(r)g(r)\quad \Rightarrow \quad
dP=-\rho(r)g(r)\,dr
\end{equation}
Putting in $g(r)$ explicitly,
\begin{equation}
\label{HTS.22} \frac{dP}{\rho(r)dr}=-\frac{Gm(r)}{r^2}
\end{equation}
Bringing the column depth back in ($y=\int \rho(r)\,dr$), this becomes
\begin{equation}
\label{HTS.23} \frac{dP}{dy}=\frac{Gm(r)}{r^2}
\end{equation}
where we've used the fact that $y=0$ at the surface and increases \emph
{inwards} (hence the sign change). Assuming $g$ is a constant, this would
give our old result
\begin{equation}
\label{HTS.24} \boxed{P=gy}
\end{equation}
Now returning to \eqref{HTS.20}:
\begin{equation}
\label{HTS.25} \frac{dP_{\mathrm{rad}}}{dy}=\frac{L(r)}{4\pi
r^2}\frac{\kappa}{c}
\end{equation}
and the result we just derived
\begin{equation}
\label{HTS.26} \frac{dP}{dy}=\frac{Gm(r)}{r^2}
\end{equation}
Taking the ratio of \eqref{HTS.25} and \eqref{HTS.26} gives us
\begin{equation}
\label{HTS.27} \frac{dP}{dP_{\mathrm{rad}}}=\frac{4\pi Gm(r)c}
{\kappa L(r)}
\end{equation}
Suppose $M=m(R)=\textrm{total mass of star}$ and $L=L(R)=\textrm
{luminosity of star}$
\begin{equation}
\label{HTS.28} \frac{dP}{dP_{\mathrm{rad}}}=\left[\frac{4\pi GcM}
{\kappa L}\right]\frac{m(r)}{M}\frac{L}{L(r)}
\end{equation}
We have now introduced the \textbf{Eddington Luminosity},
\begin{equation}
\label{HTS.29} L_{\mathrm{Edd}}=\frac{4\pi GcM}{\kappa}
\end{equation}
In general, the Eddington Luminosity is much larger than the
luminosity, since it is the luminosity where the only pressure gradient
that matters is the radiation pressure (where the star becomes unstable).
Evaluating the Eddington Luminosity in solar units and with electron
scattering,
\begin{equation}
\label{HTS.30} L_{\mathrm{Edd}}=3.1\times 10^4\ L_{\odot}\left
(\frac{M}{M_\odot}\right)
\end{equation}
Also recall how luminosity scales with mass:
\begin{equation}
\label{HTS.31} L=L_\odot\left(\frac{M}{M_\odot}\right)^3
\end{equation}
So we see that for low-mass stars, their luminosities are indeed much
smaller than the Eddington Luminosity. Now let's investigate the ratio of
the luminosity to the Eddington Luminosity:
\begin{equation}
\label{HTS.32} \frac{L}{L_{\mathrm{Edd}}}\approx 3\times 10^
{-5}\left(\frac{M}{M_\odot}\right)^2
\end{equation}
So until $M\geq 100\,M_\odot$, it will be the case that $L\ll L_
{\mathrm{Edd}}$ and thus $P_{\mathrm{rad}}\ll P$. Now returning to \eqref
{HTS.28}, we define a dimensionless quantity, $\eta(r)$ by
\begin{equation}
\label{HTS.33} \eta(r)\equiv \frac{L(r)}{L}\frac{M}{m(r)}
\end{equation}
And \eqref{HTS.28} looks like
\begin{equation}
\label{HTS.34} \frac{dP_{\mathrm{rad}}}{dP}=\frac{L}{L_{\mathrm
{Edd}}}\eta(r)
\end{equation}
Aside: this model of stars is the called the ``Eddington Standard
Model'' and was used to describe stars before their source of energy was
known.
\subsubsection{Aside: Alternate Derivation of the Eddington Luminosity} % (fold)
\label{ssub:alternate_derivation_of_the_eddington_luminosity}
Cosnider a situaton where photons are streaming through a medium of protons
and electrons with a number density of photons $n_\gamma$ with the Thomson
cross section (they interact primarily with electrons),
$\sigma_{\mathrm{es}}$. Then the collision time (the average time between
interactions with electrons) is
\begin{equation}
\label{eq:tcollision} t_{\mathrm{coll}} =
\frac{1}{\sigma_{\mathrm{es}}n_\gamma c}
\end{equation}
And the momentum change from such collisions is given simply by
\begin{equation}
\label{eq:ledd:1} \Delta p = \frac{E_\gamma}{c}=\frac{h\nu}{c}
\end{equation}
With the impulse and the time between interactions we can define a force:
\begin{equation}
\label{eq:ledd:2} F=\frac{\Delta p}{\Delta t} =
\sigma_{\mathrm{es}}n_\gamma E_\gamma = \sigma_{\mathrm{es}} \frac{F}{c}
\end{equation}
where the second $F$ is now the flux. Now when this force is greater than
local gravity, $\sigma_{\mathrm{es}} F/c > m_p g$ (now we use the proton
mass since we assume the electrons ``pull'' the protons along for the ride),
we get radiative levitation with the condition
\begin{equation}
\label{eq:ledd:3} \frac{\sigma_{\mathrm{es}}L}{4\pi r^2 c} > m_p \frac{G
M}{r^2} \qquad \Rightarrow \qquad L > L_{\mathrm{Edd}} = \frac{4\pi G M
c}{\kappa}
\end{equation}
So this is the luminosity at which the radiation field drives a mass loss
from the star. The details of what actually happens in these conditions is
still an active area of research.
% subsubsection alternate_derivation_of_the_eddington_luminosity (end)
\subsubsection{Polytropic Relations}
Returning now to the Eddinton standard model, we integrate \eqref{HTS.34} and
get
\begin{equation}
\label{HTS.35} \int_R^rdP_{\mathrm{rad}}=\frac{L}{L_{\mathrm{Edd}}}
\int_R^r\eta(r)\,dP
\end{equation}
We can integrate the left side, leaving us with
\begin{equation}
\label{HTS.36} P_{\mathrm{rad}}(r)=\frac{L}{L_{\mathrm{Edd}}}
\int_R^r\eta(r)\,dP
\end{equation}
Mathematically speaking, we are now stuck because we do not have any
knowledge of $\eta(r)$. The formal approach of solving this problem is to
define spatial averages of $\eta(r)$ for different choices of mass and
luminosity profiles. We will, however, just let $\eta(r)\approx 1$, which
gives us
\begin{equation}
\label{HTS.37} P_{\mathrm{rad}}=\frac{L}{L_{\mathrm{Edd}}}P_
{\mathrm{gas}}(r)
\end{equation}
And now substituting in our equations of state for the radiation
pressure and assuming that the radiation pressure is negligible compared to
the gas pressure:
\begin{equation}
\label{HTS.38} \frac{1}{3}aT^4=\frac{L}{L_{\mathrm{Edd}}}\frac{\rho
kT}{\mu m_p}
\end{equation}
Solving this for $T^3$, we have
\begin{equation}
\label{HTS.39} \boxed{T(r)^3=\frac{L}{L_{\mathrm{Edd}}}\frac{3k}{a
\mu m_p}\rho(r)}
\end{equation}
So we have gone from a situation where we had only typical or central
values to an actual equation that lets us find $T(r)$, $\rho(r)$, and $P(r)
$. It appears then, that in our low radiation pressure limit, $\rho\propto
T^3$. Additionally, for the ideal gas law, we see that $P\propto \rho^
{4/3}$. This is an example of the more general idea of a \textbf{polytrope}, which is a gas whose pressue and density are related simply by
\begin{equation}
\label{eq:polytrope} P\propto \rho^{1+\frac{1}{n}}
\end{equation}
Here $n$ is the \textbf{polytropic index}. These will be important later,
especially for white dwarf stars. For our $n=3$ polytrope we can take a given
$L/L_{\mathrm{Edd}}$ and recover a relation between pressure and density via
\begin{align}
\label{HTS.40} \frac{dP}{dr}&=-\rho(r)\frac{Gm(r)}{r^2}\\
\label{HTS.41} dm(r) &= \rho 4\pi r^2\,dr
\end{align}
Let us briefly consider an adiabatic change in an ideal gas. We recall
from Freshman physics that $PV^\gamma=\mathrm{const}$. For a monatomic,
ideal gas, $\gamma=5/3$, so we have $P\propto \rho^{5/3}$ and $\rho T
\propto P$, so $T\propto \rho^{2/3}$. Then we'll call the ``entropy'' $T/
\rho^{2/3}$. For our star, we have $T^3\propto \rho\Rightarrow T\propto
\rho^{1/3}$. In our case, the ``entropy'' is then $a/\rho^{1/3}$. Thus the
entropy is highest in the outer layers and smallest in the center.\\
\n At this time we should also introduce another parameter often used in
stellar astrophysics, $\beta$. Typically we designate
\begin{equation}
\label{eq:beta:1} \beta \equiv \frac{P_\mathrm{gas}}{P_{\mathrm{tot}}}
\end{equation}
This parameter gives us a sense of how radiation dominated the gas is at a
given location. We can write the radiation pressure in terms of $\beta$ and
$P_{\mathrm{gas}}$ via
\begin{equation}