Difference between revisions of "User:Tohline/SSC/Stability/Polytropes"

From VistrailsWiki
Jump to navigation Jump to search
(Delete entire subsection on n=5 polytropic spheres because this material has been moved to a new chapter)
(Delete subsection discussing Yabushita's (1991) work on truncated isothermal spheres, as it has been moved to a separate chapter on this topic)
Line 455: Line 455:
</div>
</div>
With the exception of the leading negative sign on the right-hand side, this expression is identical to the outer boundary condition identified by equation (58) of HRW66 &#8212; see the [[User:Tohline/SSC/Stability/Polytropes#HRW66excerpt|excerpt reproduced above]].
With the exception of the leading negative sign on the right-hand side, this expression is identical to the outer boundary condition identified by equation (58) of HRW66 &#8212; see the [[User:Tohline/SSC/Stability/Polytropes#HRW66excerpt|excerpt reproduced above]].
==Yabushita's (1992) Analysis==
In the portion (&sect;5) of his analysis that is focused on the stability of pressure-truncated polytropic spheres, [http://adsabs.harvard.edu/abs/1992Ap%26SS.193..173Y S. Yabushita (1992)] examined the eigenvalue problem governed by the following wave equation:
<div align="center" id="HRW66excerpt">
<table border="2" cellpadding="10">
<tr>
  <th align="center">
Radial Pulsation Equation Extracted<sup>&dagger;</sup> from p. 182 of [http://adsabs.harvard.edu/abs/1992Ap%26SS.193..173Y S. Yabushita (1992)]<p></p>
"''Similarity Between the Structure and Stability of Isothermal and Polytropic Gas Spheres''"<p></p>
Astrophysics and Space Science, vol. 193, pp. 173-183 &copy; [http://www.springer.com/astronomy/astrophysics+and+astroparticles/journal/10509 Springer]
  </th>
<tr>
  <td>
[[File:Yabushita1992WaveEquation2.png|650px|center|Yabushita (1992)]]
  </td>
</tr>
<tr><td align="left">
<sup>&dagger;</sup>Equations and text displayed here exactly as it appears in the original publication.
</td></tr>
</table>
</div>
Let's examine the overlap between this pair of governing relations and the ones employed by HRW66.  If we replace the variable <math>~X</math> with <math>~h</math>, set <math>~\gamma = (n+1)/n</math>, and set the dimensionless eigenfrequency, <math>~s</math>, to zero in the [[#HRW66excerpt|radial pulsation equation employed by HRW66]], we have,
<div align="center">
<table border="0" cellpadding="5" align="center">
<tr>
  <td align="right">
<math>~0 </math>
  </td>
  <td align="center">
<math>~=</math>
  </td>
  <td align="left">
<math>~
\frac{d^2 h}{dx^2} + \biggl[\frac{4}{x} + (n+1) \frac{\theta^'}{\theta} \biggr] \frac{dh}{dx} + (n+1)\biggl[ 3 - \frac{4n}{(n+1)} \biggr] \biggl[ \frac{\theta^' h}{\theta x} \biggr]
</math>
  </td>
</tr>
<tr>
  <td align="right">
&nbsp;
  </td>
  <td align="center">
<math>~=</math>
  </td>
  <td align="left">
<math>~
\frac{d^2 h}{dx^2} + \biggl[\frac{4}{x} + (n+1) \frac{\theta^'}{\theta} \biggr] \frac{dh}{dx} + (3-n) \biggl[ \frac{\theta^' h}{\theta x} \biggr] \, .
</math>
  </td>
</tr>
</table>
</div>
This matches equation (5.3) of [http://adsabs.harvard.edu/abs/1992Ap%26SS.193..173Y Yabushita (1992)] &#8212; see the above boxed-in image &#8212; except the <math>~(4/x)</math> term appears as <math>~(2/x)</math> in Yabushita's article; giving the benefit of the doubt, <font color="red">this is most likely a typographical error</font> in [http://adsabs.harvard.edu/abs/1992Ap%26SS.193..173Y Yabushita (1992)].  According to HRW66, the corresponding central boundary condition is,
<div align="center">
<math>\frac{dh}{dx} = 0</math> &nbsp; &nbsp; &nbsp; &nbsp; at &nbsp; &nbsp; &nbsp; &nbsp; <math>x=0 \, .</math>
</div>
While &#8212; after changing the sign on the right-hand side of HRW66's equation (58) as argued in our [[User:Tohline/SSC/Perturbations#ChristyCox|accompanying discussion]] in order to align with the separate derivations presented by [http://adsabs.harvard.edu/abs/1966ARA%26A...4..353C Christy (1965)] and [http://adsabs.harvard.edu/abs/1967IAUS...28....3C Cox (1967)]  &#8212; the corresponding boundary condition at the surface is,
<div align="center">
<table border="0" cellpadding="5" align="center">
<tr>
  <td align="right">
<math>~\frac{dh}{dx}</math>
  </td>
  <td align="center">
<math>~=</math>
  </td>
  <td align="left">
<math>~- \frac{h}{x} \biggr[  3 - \frac{4}{\gamma}  + \cancelto{0}{\frac{x s^2}{\gamma q}} \biggr]</math>
  </td>
  <td align="left" colspan="2">
&nbsp;
  </td>
</tr>
<tr>
  <td align="right">
&nbsp;
  </td>
  <td align="center">
<math>~=</math>
  </td>
  <td align="left">
<math>~\frac{n-3}{n+1} \biggl(\frac{h}{x} \biggr) \, .</math>
  </td>
  <td align="left">
&nbsp; &nbsp; &nbsp; &nbsp; at &nbsp; &nbsp; &nbsp; &nbsp;
  </td>
  <td align="left">
<math>~x = x_0 \, .</math>
  </td>
</tr>
</table>
</div>
This surface boundary condition, which has been used by the astrophysics community in the context of ''isolated'' polytropic configurations, is different from the one displayed as equation (5.4) of [http://adsabs.harvard.edu/abs/1992Ap%26SS.193..173Y Yabushita (1992)].  The surface boundary condition chosen by Yabushita &#8212; effectively,
<div align="center">
<math>~\frac{d \ln h}{d\ln x} = -3 \, ,</math>
</div>
&#8212; does seem to be more appropriate in the context of a study of the stability of ''pressure-truncated'' polytropes because, as argued by [http://adsabs.harvard.edu/abs/1941ApJ....94..124L Ledoux &amp; Pekeris (1941)] and as reviewed in our [[User:Tohline/SSC/Perturbations#Set_the_Surface_Pressure_Fluctuation_to_Zero|accompanying discussion]], it ensures that the pressure fluctuation ''at the surface'' is zero.  It is worth noting that Yabushita's surface boundary condition matches the surface boundary condition chosen by [http://adsabs.harvard.edu/abs/1974MNRAS.168..427T Taff &amp; Van Horn (1974)] in their study of pressure-truncated ''isothermal'' spheres; in their words (see p. 428 of their article):  &nbsp; [Setting the surface logarithmic derivative to negative 3] <font color="green">expresses the condition that the pressure at the perturbed surface always remain[s] equal to the confining pressure exerted by the external medium in which the [pressure-truncated] sphere must be embedded</font>.


==Overview==
==Overview==

Revision as of 21:00, 16 February 2017

Radial Oscillations of Polytropic Spheres

Whitworth's (1981) Isothermal Free-Energy Surface
|   Tiled Menu   |   Tables of Content   |  Banner Video   |  Tohline Home Page   |

Groundwork

Adiabatic (Polytropic) Wave Equation

In an accompanying discussion, we derived the so-called,

Adiabatic Wave (or Radial Pulsation) Equation

LSU Key.png

<math>~ \frac{d^2x}{dr_0^2} + \biggl[\frac{4}{r_0} - \biggl(\frac{g_0 \rho_0}{P_0}\biggr) \biggr] \frac{dx}{dr_0} + \biggl(\frac{\rho_0}{\gamma_\mathrm{g} P_0} \biggr)\biggl[\omega^2 + (4 - 3\gamma_\mathrm{g})\frac{g_0}{r_0} \biggr] x = 0 </math>


whose solution gives eigenfunctions that describe various radial modes of oscillation in spherically symmetric, self-gravitating fluid configurations. If the initial, unperturbed equilibrium configuration is a polytropic sphere whose internal structure is defined by the function, <math>~\theta(\xi)</math>, then

<math>~r_0</math>

<math>~=</math>

<math>~a_n \xi \, ,</math>

<math>~\rho_0</math>

<math>~=</math>

<math>~\rho_c \theta^{n} \, ,</math>

<math>~P_0</math>

<math>~=</math>

<math>~K\rho_0^{(n+1)/n} = K\rho_c^{(n+1)/n} \theta^{n+1} \, ,</math>

<math>~g_0</math>

<math>~=</math>

<math>~\frac{GM(r_0)}{r_0^2} = \frac{G}{r_0^2} \biggl[ 4\pi a_n^3 \rho_c \biggl(-\xi^2 \frac{d\theta}{d\xi}\biggr) \biggr] \, ,</math>

where,

<math>~a_n</math>

<math>~=</math>

<math>~\biggl[\frac{(n+1)K}{4\pi G} \cdot \rho_c^{(1-n)/n} \biggr]^{1/2} \, .</math>

Hence, after multiplying through by <math>~a_n^2</math>, the above adiabatic wave equation can be rewritten in the form,

<math>~\frac{d^2x}{d\xi^2} + \biggl[\frac{4}{\xi} - \frac{g_0}{a_n}\biggl(\frac{a_n^2 \rho_0}{P_0}\biggr) \biggr] \frac{dx}{d\xi} + \biggl(\frac{a_n^2\rho_0}{\gamma_\mathrm{g} P_0} \biggr)\biggl[\omega^2 + (4 - 3\gamma_\mathrm{g})\frac{g_0}{a_n\xi} \biggr] x </math>

<math>~=</math>

<math>~0 \, .</math>

In addition, given that,

<math>~\frac{g_0}{a_n}</math>

<math>~=</math>

<math>~4\pi G \rho_c \biggl(-\frac{d \theta}{d\xi} \biggr) \, ,</math>

and,

<math>~\frac{a_n^2 \rho_0}{P_0}</math>

<math>~=</math>

<math>~\frac{(n+1)}{(4\pi G\rho_c)\theta} = \frac{a_n^2 \rho_c}{P_c} \cdot \frac{\theta_c}{\theta}\, ,</math>

we can write,

<math>~\frac{d^2x}{d\xi^2} + \biggl[\frac{4 - (n+1)V(\xi)}{\xi} \biggr] \frac{dx}{d\xi} + \biggl[\omega^2 \biggl(\frac{a_n^2 \rho_c }{\gamma_g P_c} \biggr) \frac{\theta_c}{\theta} - \biggl(3-\frac{4}{\gamma_g}\biggr) \cdot \frac{(n+1)V(x)}{\xi^2} \biggr] x </math>

<math>~=</math>

<math>0 \, ,</math>

where we have adopted the function notation,

<math>~V(\xi)</math>

<math>~\equiv</math>

<math>~- \frac{\xi}{\theta} \frac{d \theta}{d\xi} \, .</math>

As can be seen in the following framed image, this is the form of the polytropic wave equation published by J. O. Murphy & R. Fiedler (1985b, Proc. Astron. Soc. Australia, 6, 222), at the beginning of their discussion of "Radial Pulsations and Vibrational Stability of a Sequence of Two Zone Polytropic Stellar Models." (NOTE: There appears to be a sign error in the numerator of the second term of their published expression; there also appears to be an error in the definition of the coefficient, <math>~\alpha^*</math>, as given in the text of their paper.)

Polytropic Wave Equation extracted from J. O. Murphy & R. Fiedler (1985b)

"Radial Pulsations and Vibrational Stability of a Sequence of Two Zone Polytropic Stellar Models"

Proceeding of the Astronomical Society of Australia, vol. 6, pp. 222 - 226 © Astronomical Society of Australia

Murphy & Fiedler (1985b)
Equations displayed here, as a single digital image, with layout modified from the original publication.


It is also the same as the radial pulsation equation for polytropic configurations that appears as equation (56) in the detailed discussion of "The Oscillations of Gas Spheres" published by H M. Hurley, P. H. Roberts, & K. Wright (1966, ApJ, 143, 535); hereafter, we will refer to this paper as HRW66. The relevant set of equations from HRW66 has been extracted as a single digital image and reprinted, here, as a boxed-in image.

Radial Pulsation Equation as Presented by M. Hurley, P. H. Roberts, & K. Wright (1966)

"The Oscillations of Gas Spheres"

The Astrophysical Journal, vol. 143, pp. 535 - 551 © American Astronomical Society

Hurley, Roberts & Wright (1966)

Set of equations and accompanying text displayed here, as a single digital image, exactly as they appear in the original publication.

In order to make clearer the correspondence between our derived expression and the one published by HRW66, we will rewrite the HRW66 radial pulsation equation: (1) Gathering all terms on the same side of the equation; (2) making the substitution,

<math>\theta^' \rightarrow -\frac{\theta V}{x} \, ;</math>

and (3) reattaching a "prime" to the quantity, <math>~s</math>, to emphasize that it is a dimensionless frequency.

ASIDE: In their equation (46), HRW66 convert the eigenfrequency, <math>~s</math> — which has units of inverse time — to a dimensionless eigenfrequency, <math>~s^'</math>, via the relation,

<math>~s = \biggl( \frac{4\pi G \rho_c}{1+n} \biggr)^{1/2} s^' ~~~~~~~\cdots\cdots~~~~~~~(46)</math>

Then, immediately following equation (46), they state that they will "omit the prime on <math>~s</math> henceforward." As a result, the dimensionless eigenfrequency that appears in their equations (56) and (58) is unprimed. This is unfortunate as it somewhat muddies our efforts, here, to demonstrate the correspondence between the HRW66 polytropic radial pulsation equation and ours. In our subsequent manipulation of equation (56) from HRW66 we reattach a prime to the quantity, <math>~s</math>, to emphasize that it is a dimensionless frequency. But this prime on <math>~s</math> should not be confused with the prime on <math>~\theta</math> (HRW66 equation 56) or with the prime on <math>~X</math> (HRW66 equation 57), both of which denote differentiation with respect to the radial coordinate.

With these modifications, the HRW66 radial pulsation equation becomes,

<math>~0</math>

<math>~=</math>

<math> ~\frac{d^2 X}{dx^2} + \biggl[\frac{4 - (n+1)V }{x}\biggr]\frac{dX}{dx} - \frac{V}{\gamma x^2}\biggl[\frac{x^2 (s^')^2}{\theta V} + (3\gamma -4)(n+1) \biggr]X </math>

 

<math>~=</math>

<math> ~\frac{d^2 X}{dx^2} + \biggl[\frac{4 - (n+1)V }{x}\biggr]\frac{dX}{dx} + \biggl[-\frac{(s^')^2 }{\gamma \theta } - \biggl(3 -\frac{4}{\gamma}\biggr)\frac{(n+1)V}{x^2} \biggr]X \, . </math>

The correspondence with our derived expression is complete, assuming that,

<math>~(s^')^2</math>

<math>~=</math>

<math>~-\omega^2 \biggl(\frac{a_n^2 \rho_c }{P_c} \biggr) \theta_c</math>

 

<math>~=</math>

<math>~-\omega^2 \biggl[\frac{n+1 }{4\pi G \rho_c} \biggr] \, .</math>

As has been explained in the above "ASIDE," this is exactly the factor that HRW66 use to normalize their eigenfrequency, <math>~s</math>, and make it dimensionless <math>~(s^')</math>. It is clear, as well, that HRW66 have adopted a sign convention for the square of their eigenfrequency that is the opposite of the sign convention that we have adopted for <math>~\omega^2</math>. That is, it is clear that,

<math>~s^2 ~~\leftrightarrow~~ - \omega^2 \, .</math>

Boundary Conditions

As we have pointed out in the context of a general discussion of boundary conditions associated with the adiabatic wave equation, the eigenfunction, <math>~x</math>, will be suitably well behaved at the center of the configuration if,

<math>~\frac{dx}{dr_0} = 0</math>        at         <math>~r_0 = 0 \, ,</math>

which, in the context of our present discussion of polytropic configurations, leads to the inner boundary condition,

<math>~\frac{dx}{d\xi} = 0</math>        at         <math>~\xi = 0 \, .</math>

This is precisely the inner boundary condition specified by HRW66 — see their equation (57), which has been reproduced in the above excerpt from HWR66.


As we have also shown in the context of this separate, general discussion of boundary conditions associated with the adiabatic wave equation, the pressure fluctuation will be finite at the surface — even if the equilibrium pressure and/or the pressure scale height go to zero at the surface — if the radial eigenfunction, <math>~x</math>, obeys the relation,

<math>~r_0 \frac{dx}{dr_0}</math>

<math>~=</math>

<math>~\biggl( 4 - 3\gamma_g + \frac{\omega^2 R^3}{GM_\mathrm{tot}}\biggr) \frac{x}{\gamma_g}</math>        at         <math>~r_0 = R \, .</math>

Or, given that, in polytropic configurations, <math>~r_0 = a_n\xi</math>,

<math>~\xi \frac{dx}{d\xi}</math>

<math>~=</math>

<math>~\frac{x}{\gamma_g} \biggl[ 4 - 3\gamma_g + \frac{\omega^2 (a_n \xi_1)^3}{GM_\mathrm{tot}}\biggr] </math>        at         <math>~\xi = \xi_1 \, ,</math>

where, the subscript "1" denotes equilibrium, surface values. As can be deduced from our above summary of the properties of polytropic configurations,

<math>~GM_\mathrm{tot}</math>

<math>~=</math>

<math>~4\pi G a_n^3 \rho_c (-\xi_1^2 \theta_1^') \, .</math>

Hence, for spherically symmetric polytropic configurations, the surface boundary condition becomes,

<math>~\frac{dx}{d\xi}</math>

<math>~=</math>

<math>~\frac{x}{\gamma_g \xi} \biggl[ 4 - 3\gamma_g + \omega^2 \biggl( \frac{1}{4\pi G \rho_c } \biggr) \frac{\xi}{(-\theta^')}\biggr] </math>         at         <math>~\xi = \xi_1 \, ,</math>

<math>~\Rightarrow ~~~~~(n+1)\frac{dx}{d\xi}</math>

<math>~=</math>

<math>~\frac{x}{\gamma_g \xi} \biggl[ (n+1)(4 - 3\gamma_g) + \omega^2 \biggl( \frac{1+n}{4\pi G \rho_c } \biggr) \frac{\xi}{(-\theta^')}\biggr] </math>

 

<math>~=</math>

<math>~-\frac{x}{\gamma_g \xi} \biggl[ (n+1)(3\gamma_g-4) - \omega^2 \biggl( \frac{1+n}{4\pi G \rho_c } \biggr) \frac{\xi}{(-\theta^')}\biggr] </math>         at         <math>~\xi = \xi_1 \, .</math>

Adopting notation used by HRW66, specifically, as demonstrated above,

<math>~-\omega^2 \biggl( \frac{1+n}{4\pi G \rho_c } \biggr) \rightarrow (s^')^2 \, , </math>

and, from equation (50) of HRW66,

<math>~-\theta^' \rightarrow q </math>         at         <math>~\xi = \xi_1 \, ,</math>

this outer boundary condition becomes,

<math>~(n+1)\frac{dx}{d\xi}</math>

<math>~=</math>

<math>~-\frac{x}{\gamma_g \xi} \biggl[ (n+1)(3\gamma_g-4) + \frac{\xi (s^')^2}{q}\biggr] </math>         at         <math>~\xi = \xi_1 \, .</math>

With the exception of the leading negative sign on the right-hand side, this expression is identical to the outer boundary condition identified by equation (58) of HRW66 — see the excerpt reproduced above.

Overview

The eigenvector associated with radial oscillations in isolated polytropes has been determined numerically and the results have been presented in a variety of key publications:

Tables

Quantitative Information Regarding Eigenvectors of Oscillating Polytropes

<math>~(\Gamma_1 = 5/3)</math>

<math>~n</math>

<math>~\frac{\rho_c}{\bar\rho}</math>

Excerpts from Table 1 of

Hurley, Roberts, & Wright (1966)

<math>~s^2 (n+1)/(4\pi G\rho_c)</math>

Excerpts from Table 3 of

J. P. Cox (1974)

<math>~\sigma_0^2 R^3/(GM)</math>

<math>\frac{(n+1) *\mathrm{Cox74}}{3 *\mathrm{HRW66}} \cdot \frac{\bar\rho}{\rho_c}</math>

<math>~0</math>

<math>~1</math>

<math>~1/3</math>

<math>~1</math>

<math>~1</math>

<math>~1</math>

<math>~3.30</math>

<math>~0.38331</math>

<math>~1.892</math>

<math>~0.997</math>

<math>~1.5</math>

<math>~5.99</math>

<math>~0.37640</math>

<math>~2.712</math>

<math>~1.002</math>

<math>~2</math>

<math>~11.4</math>

<math>~0.35087</math>

<math>~4.00</math>

<math>~1.000</math>

<math>~3</math>

<math>~54.2</math>

<math>~0.22774</math>

<math>~9.261</math>

<math>~1.000</math>

<math>~3.5</math>

<math>~153</math>

<math>~0.12404</math>

<math>~12.69</math>

<math>~1.003</math>

<math>~4.0</math>

<math>~632</math>

<math>~0.04056</math>

<math>~15.38</math>

<math>~1.000</math>


n = 3 Polytrope

Here we perform a numerical integration of the governing LAWE for <math>~n=3</math> polytropes. We can directly compare our results with Schwarzschild's (1941) published work on "Overtone Pulsations for the Standard [Stellar] Model."

Drawing from our above discussion, the LAWE for any polytrope of index, <math>~n</math>, may be written as,

<math>~0 </math>

<math>~=</math>

<math>~\frac{d^2x}{d\xi^2} + \biggl[\frac{4 - (n+1)V(\xi)}{\xi} \biggr] \frac{dx}{d\xi} + \biggl[\omega^2 \biggl(\frac{a_n^2 \rho_c }{\gamma_g P_c} \biggr) \frac{\theta_c}{\theta} - \biggl(3-\frac{4}{\gamma_g}\biggr) \cdot \frac{(n+1)V(x)}{\xi^2} \biggr] x </math>

 

<math>~=</math>

<math>~\frac{d^2x}{d\xi^2} + \biggl[\frac{4}{\xi} - \frac{(n+1)}{\theta} \biggl(- \frac{d\theta}{d\xi} \biggr)\biggr] \frac{dx}{d\xi} + \frac{(n+1)}{\theta} \biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi } \biggl(- \frac{d\theta}{d\xi} \biggr) \biggr] x </math>

where,

<math>~\sigma_c^2</math>

<math>~\equiv</math>

<math>~\frac{3\omega^2}{2\pi G\rho_c} \, .</math>

It can be shown straightforwardly that this matches the LAWE used by Schwarzschild (1941), if <math>~n</math> is set to 3. But let's postpone making this substitution until we formulate a general approach to integrating this equation from the center of the configuration, outward. Following a parallel discussion, we begin by multiplying the LAWE through by <math>~(\xi\theta)</math>, obtaining a 2nd-order ODE that is relevant at every individual coordinate location, <math>~\xi_i</math>, namely,

<math>~\theta_i {x_i}</math>

<math>~=</math>

<math>~- \biggl[4\theta_i - (n+1)\xi_i (- \theta^')_i\biggr] \frac{x_i'}{\xi_i} - (n+1)\biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi_i } (- \theta^')_i\biggr] x_i </math>

Now, using the general finite-difference approach described separately, we make the substitutions,

<math>~x_i'</math>

<math>~\approx</math>

<math>~ \frac{x_+ - x_-}{2 \Delta_\xi} \, ; </math>

and,

<math>~ x_i </math>

<math>~\approx</math>

<math>~\frac{x_+ - 2x_i + x_-}{\Delta_\xi^2} \, ,</math>

which will provide an approximate expression for <math>~x_+ \equiv x_{i+1}</math>, given the values of <math>~x_- \equiv x_{i-1}</math> and <math>~x_i</math>. Specifically, if the center of the configuration is denoted by the grid index, <math>~i=1</math>, then for zones, <math>~i = 3 \rightarrow N</math>,

<math>~\theta_i \biggl[ \frac{x_+ - 2x_i + x_-}{\Delta_\xi^2} \biggr]</math>

<math>~=</math>

<math>~- \biggl[4\theta_i - (n+1)\xi_i (- \theta^')_i\biggr] \biggl[ \frac{x_+ - x_-}{2 \xi_i \Delta_\xi} \biggr] - (n+1)\biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi_i } (- \theta^')_i\biggr] x_i </math>

<math>~\Rightarrow ~~~ \theta_i \biggl[ \frac{x_+ }{\Delta_\xi^2} \biggr] + \biggl[4\theta_i - (n+1)\xi_i (- \theta^')_i\biggr] \biggl[ \frac{x_+ }{2 \xi_i\Delta_\xi} \biggr]</math>

<math>~=</math>

<math>~ -\theta_i \biggl[ \frac{- 2x_i + x_-}{\Delta_\xi^2} \biggr] - \biggl[4\theta_i - (n+1)\xi_i (- \theta^')_i\biggr] \biggl[ \frac{- x_-}{2 \xi_i \Delta_\xi} \biggr] - (n+1)\biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi_i } (- \theta^')_i\biggr] x_i </math>

<math>~\Rightarrow ~~~ x_+ \biggl[2\theta_i +\frac{4\Delta_\xi \theta_i}{\xi_i} - \Delta_\xi (n+1)(- \theta^')_i\biggr] </math>

<math>~=</math>

<math>~ x_- \biggl[\frac{4\Delta_\xi \theta_i}{\xi_i} - \Delta_\xi (n+1)(- \theta^')_i - 2\theta_i\biggr] + x_i\biggl\{4\theta_i - 2\Delta_\xi^2(n+1)\biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi_i } (- \theta^')_i\biggr] \biggr\} </math>

 

<math>~=</math>

<math>~ x_- \biggl[\frac{4\Delta_\xi \theta_i}{\xi_i} - \Delta_\xi (n+1)(- \theta^')_i - 2\theta_i\biggr] + x_i\biggl\{4\theta_i - \frac{\Delta_\xi^2(n+1)}{3}\biggl[ \frac{\sigma_c^2}{\gamma_g} - 2\alpha \biggl(- \frac{3\theta^'}{\xi}\biggr)_i\biggr] \biggr\} \, .</math>

In order to kick-start the integration, at the center of the configuration <math>~(\xi_1 = 0)</math>, we will set the eigenfunction value to <math>~x_1 = 1</math>; and we will use the assumed symmetry condition, <math>~x_1' = 0 ~~\Rightarrow ~~ x_- = x_+</math>, in order to determine the value of the eigenfunction at the first grid location off center <math>~(\xi_2 = \Delta)</math>. That is, for <math>~i = 1</math>, the discretized LAWE gives,

<math>~ x_+ </math>

<math>~=</math>

<math>~ x_i \biggl\{ 1 -\frac{\Delta_\xi^2 (n+1)}{2\theta_i}\biggl[ \frac{\sigma_c^2}{6\gamma_g} - \frac{\alpha}{\xi_i } (- \theta^')_i\biggr] \biggr\} \, .</math>

Given that the power-series representation of the Lane-Emden function is,

<math>~\theta</math>

<math>~=</math>

<math>~ 1 - \frac{\xi^2}{6} + \frac{n}{120} \xi^4 - \frac{n}{378} \biggl( \frac{n}{5} - \frac{1}{8} \biggr) \xi^6 + \biggl[ \frac{n(122n^2 -183n + 70)}{3265920} \biggr] \xi^8 + \cdots </math>

we know that in this individual case <math>~(i=1)</math> — that is, for <math>~\xi = 0</math> — <math>~\theta = 1</math> and <math>~(-\theta')/\xi = \tfrac{1}{3}</math> for all values of the polytropic index. Hence, for <math>~i=1</math>, we can use the expression,

<math>~ x_2 </math>

<math>~=</math>

<math>~ x_1 [ 1 - \tfrac{1}{12} \Delta_\xi^2 (n+1) \mathfrak{F} ] \, ,</math>

where,

<math>~ \mathfrak{F} </math>

<math>~\equiv</math>

<math>~ \biggl[ \frac{\sigma_c^2}{\gamma_g} - 2\alpha\biggr]\, .</math>

Related Discussions

  • In an accompanying Chapter within our "Ramblings" Appendix, we have played with the adiabatic wave equation for polytropes, examining its form when the primary perturbation variable is an enthalpy-like quantity, rather than the radial displacement of a spherical mass shell. This was done in an effort to mimic the approach that has been taken in studies of the stability of Papaloizou-Pringle tori.
  • <math>~n=3</math> … M. Schwarzschild (1941, ApJ, 94, 245), Overtone Pulsations of the Standard Model: This work is referenced in §38.3 of [KW94]. It contains an analysis of the radial modes of oscillation of <math>~n=3</math> polytropes, assuming various values of the adiabatic exponent.
  • <math>~n=\tfrac{3}{2}</math> … D. Lucas (1953, Bul. Soc. Roy. Sci. Liege, 25, 585) … Citation obtained from the Prasad & Gurm (1961) article.
  • <math>~n=1</math> … L. D. Chatterji (1951, Proc. Nat. Inst. Sci. [India], 17, 467) … Citation obtained from the Prasad & Gurm (1961) article.


Whitworth's (1981) Isothermal Free-Energy Surface

© 2014 - 2021 by Joel E. Tohline
|   H_Book Home   |   YouTube   |
Appendices: | Equations | Variables | References | Ramblings | Images | myphys.lsu | ADS |
Recommended citation:   Tohline, Joel E. (2021), The Structure, Stability, & Dynamics of Self-Gravitating Fluids, a (MediaWiki-based) Vistrails.org publication, https://www.vistrails.org/index.php/User:Tohline/citation