dp dr = ρgm r 2 (1) dm dt dr = 3

Similar documents
Astronomy 112: The Physics of Stars. Class 9 Notes: Polytropes

Summary of stellar equations

dr + Gm dr Gm2 2πr 5 = Gm2 2πr 5 < 0 (2) Q(r 0) P c, Q(r R) GM 2 /8πR 4. M and R are total mass and radius. Central pressure P c (Milne inequality):

ASTM109 Stellar Structure and Evolution Duration: 2.5 hours

Astronomy 112: The Physics of Stars. Class 13 Notes: Schematics of the Evolution of Stellar Cores

SPA7023P/SPA7023U/ASTM109 Stellar Structure and Evolution Duration: 2.5 hours

Week 8: Stellar winds So far, we have been discussing stars as though they have constant masses throughout their lifetimes. On the other hand, toward

Physics 576 Stellar Astrophysics Prof. James Buckley. Lecture 11 Polytropes and the Lane Emden Equation

Introduction to Astrophysics Tutorial 2: Polytropic Models

Astronomy 112: The Physics of Stars. Class 14 Notes: The Main Sequence

THIRD-YEAR ASTROPHYSICS

Ay 1 Lecture 8. Stellar Structure and the Sun

9.1 Introduction. 9.2 Static Models STELLAR MODELS

CHAPTER 16. Hydrostatic Equilibrium & Stellar Structure

3 Hydrostatic Equilibrium

P M 2 R 4. (3) To determine the luminosity, we now turn to the radiative diffusion equation,

dp dr = GM c = κl 4πcr 2

PAPER 58 STRUCTURE AND EVOLUTION OF STARS

How the mass relationship with a White Dwarf Star

Static Stellar Structure

2. Equations of Stellar Structure

Physics 556 Stellar Astrophysics Prof. James Buckley

TRANSFER OF RADIATION

Ordinary Differential Equations

Stellar Structure and Evolution

7. The Evolution of Stars a schematic picture (Heavily inspired on Chapter 7 of Prialnik)

Physics 576 Stellar Astrophysics Prof. James Buckley. Lecture 10

Stellar Interiors - Hydrostatic Equilibrium and Ignition on the Main Sequence.

ρ dr = GM r r 2 Let s assume that µ(r) is known, either as an arbitrary input state or on the basis of pre-computed evolutionary history.

Stellar Models ASTR 2110 Sarazin

Fundamental Stellar Parameters. Radiative Transfer. Stellar Atmospheres

Lecture 1. Overview Time Scales, Temperature-density Scalings, Critical Masses

Lecture 1. Overview Time Scales, Temperature-density Scalings, Critical Masses. I. Preliminaries

Stars Lab 2: Polytropic Stellar Models

Stellar Evolution ASTR 2110 Sarazin. HR Diagram vs. Mass

Chapter 7 Neutron Stars

Polytropic Stars. c 2

Stellar Spectra ASTR 2120 Sarazin. Solar Spectrum

Introduction. Stellar Objects: Introduction 1. Why should we care about star astrophysics?

The Stellar Opacity. F ν = D U = 1 3 vl n = 1 3. and that, when integrated over all energies,

mc 2, (8.1) = R Sch 2R

Star Formation and Protostars

Phases of Stellar Evolution

AST1100 Lecture Notes

PHY-105: Nuclear Reactions in Stars (continued)

Before proceeding to Chapter 20 More on Cluster H-R diagrams: The key to the chronology of our Galaxy Below are two important HR diagrams:

Astro 1050 Wed. Apr. 5, 2017

AST1100 Lecture Notes

M J v2 s. Gm (r) P dm = 3. For an isothermal gas, using the fact that m r, this is GM 2 R = 3N ot M 4πR 3 P o, P o = 3N ot M

Astronomy II (ASTR1020) Exam 3 Test No. 3D

Stellar Winds: Mechanisms and Dynamics

(2) low-mass stars: ideal-gas law, Kramer s opacity law, i.e. T THE STRUCTURE OF MAIN-SEQUENCE STARS (ZG: 16.2; CO 10.6, 13.

Equations of Stellar Structure

Astro 1050 Fri. Apr. 10, 2015

Lecture 7.1: Pulsating Stars

1.1 An introduction to the cgs units

The Interiors of the Stars

Linear Theory of Stellar Pulsation

Life and Death of a Star. Chapters 20 and 21

Using that density as the electronic density, we find the following table of information for the requested quantities:

Astronomy 112: The Physics of Stars. Class 5 Notes: The Pressure of Stellar Material

read 9.4-end 9.8(HW#6), 9.9(HW#7), 9.11(HW#8) We are proceding to Chap 10 stellar old age

Substellar Interiors. PHY 688, Lecture 13

11/19/08. Gravitational equilibrium: The outward push of pressure balances the inward pull of gravity. Weight of upper layers compresses lower layers

Stellar energy generation on the main sequence

The Sun. Nearest Star Contains most of the mass of the solar system Source of heat and illumination

Assignment 4 Solutions [Revision : 1.4]

Overview of Astronomical Concepts III. Stellar Atmospheres; Spectroscopy. PHY 688, Lecture 5 Stanimir Metchev

Convection. If luminosity is transported by radiation, then it must obey

The structure and evolution of stars

Lecture 7: Stellar evolution I: Low-mass stars

t KH = GM2 RL Pressure Supported Core for a Massive Star Consider a dense core supported by pressure. This core must satisfy the equation:

Ay123 Set 3 solutions

7/9. What happens to a star depends almost completely on the mass of the star. Mass Categories: Low-Mass Stars 0.2 solar masses and less

ASTR3007/4007/6007, Class 2: Stellar Masses; the Virial Theorem

The Equations of Stellar structure

Summary of Results. 1. Mass Conserva-on. dr dm = 1. dm dr = 4πr 2 ρ. 4πr 2 ρ

Lec 9: Stellar Evolution and DeathBirth and. Why do stars leave main sequence? What conditions are required for elements. Text

Physics 556 Stellar Astrophysics Prof. James Buckley. Lecture 9 Energy Production and Scaling Laws

Nuclear Reactions and Solar Neutrinos ASTR 2110 Sarazin. Davis Solar Neutrino Experiment

Lecture 7. Evolution of Massive Stars on the Main Sequence and During Helium Burning - Basics

Model Atmospheres. Model Atmosphere Assumptions

Opacity and Optical Depth

V. Stars.

The Stellar Black Hole

Chapter 16. White Dwarfs

Tides in Higher-Dimensional Newtonian Gravity

UNIVERSITY OF SOUTHAMPTON

Core Collapse Supernovae

PHAS3135 The Physics of Stars

Atoms and Star Formation

7 Stellar Winds & Supernovae Remnants

Stellar evolution in a nutshell

class 21 Astro 16: Astrophysics: Stars, ISM, Galaxies November 20, 2018

10/17/2012. Stellar Evolution. Lecture 14. NGC 7635: The Bubble Nebula (APOD) Prelim Results. Mean = 75.7 Stdev = 14.7

Order of Magnitude Estimates in Quantum

Mass-Radius Relation: Hydrogen Burning Stars

Energy transport: convection

3. Stellar radial pulsation and stability

Ay123 Set 1 solutions

Transcription:

Week 5: Simple equilibrium configurations for stars Follows on closely from chapter 5 of Prialnik. Having understood gas, radiation and nuclear physics at the level necessary to have an understanding of the basic processes inside stars, we can now move to developing some formalisms for the basic equilibrium configurations of stars. There are four basic equations that need to be solved. Here we will have already made one simplifying assumption, that the chemical composition of the star is constant throughout this is a good approximation for getting a basic picture of what the main sequence structure of a star looks like, but obviously if this approximation is taken too far, then the star cannot evolve. dr = ρgm r 2 1) dm dr = 4πr2 ρ 2) dt dr = 3 4ac κρ T 3 F 4πr 2 3) df dr = 4πr2 ρq 4) To these, we add a few relations that describe the pressure, opacity, and heat generation: P = R µ I ρt + P e + 1 3 at 4 5) κ = κ 0 ρ a T b 6) q = q 0 ρ m T n 7) Additionally, since we have four first order differential equations, we need four boundary conditions. Three are obvious: at the center, m = 0 and F = 0, and at the surface, P = 0. Note that this one is only an approximation, depending on how the surface is defined.) The fourth boundary condition must be some relation between the temperature at the surface and the temperature somewhere in the stellar interior. One can see that this will not be an easy set of equations to solve, and that a numerical approach will be required. The equations are coupled and nonlinear, and the value of pressure varies by 10 11 between the surface and center, and the value of temperature varies by more than a factor of 1000. Despite all this, considerable progress was made on understanding stellar structure long before 1

the development of electronic computers 1, and before the discovery of the key bits of nuclear physics that allow us to understand energy generation in stars. What allowed progress to be made was the development of some key simplifications which turn out to be reasonable approximations of what really happens in stars. We have already discussed one the assumption of uniform composition. 1 Polytropes One simplifying assumption that can be made is that the equation of state in the star is constant. This is sometimes a very good assumption, and sometimes poor but it allows for a straightforward calculation of the structure of something like a star. We can start with the equation 1, multiply by r 2 /ρ, and differentiate with respect to r: ) d r 2 = G dm dr ρ dr dr, 8) and we get an equation whose right side can be substituted with 4πGρ. Now, we set P = Kρ γ, a polytropic equation of state, defined by K and n, the polytropic index, which is defined such that γ = 1 + 1 n. Then, substituting in for P with Kρ 1+ 1 1 everywhere, we get: n + 1)K 4πGn 1 r 2 d dr r 2 ρ n 1 n We then set up several substitutions to non-dimensionalize the equations: dρ dr ) 9) ρ = ρ c θ n 10) where ρ c is the central density; [ ] n + 1)K = α 2 11) 4πGρ n 1 n c and r = αξ 12) which we can then combine to produce the Lane-Emden equation: 1 d ξ 2 ξ 2 dθ ) = θ n. 13) dξ dξ 1 Edward Charles Pickering, director of the Harvard Observatory around the turn of the 20th century, had a group of women who worked under him who did calculations by hand, and also classifying spectra of stars, and were known as the Harvard Computers. A few, notably Annie Jump Cannon, made major contributions to astronomy that went well beyond just doing calculations and classifying spectra. 2

This equation has two boundary conditions: θ = 1 and dθ/dξ = 0 at ξ = 0. This still requires numerical integration unless n = 0 or n = 1, but the numerical integration is fairly straightforward. When ξ = 0, the value is ξ is referred to as ξ 1, and the surface of the star has been reached: R = αξ 1. The mass of the polytrope is given as: M = R 0 dθ dξ ξ1 4πr 2 ρdr = 4πα 3 ρ c ξ 2 θ n dξ 14) ) Then, M = 4πα 3 ρ c ξ1 2, based on substituting in θ n from the original ξ 1 Lane-Emden equation. For n = 0 and n = 1, there are simple analytic solutions to the Lane-Emden equation, but in general, numerical calculations are needed but, unlike for the case of the full set of stellar evolution equations, these numerical calculations can be done in a straightforward way, since we now have transformed the problem into integration of a simple second order differential equation in one variable. We can define 4 polytropic constants D n, M N, R N, and B n, where the first three are easy to remember, since they relate to density, mass, and radius, and the fourth is a number close to 1 that serves as a catch-all in the final pressure equation. ρ c = D n ρ 15) [ )] 1 3 dθ defines D n, which is equal to ξ 1 dξ. Next, we can take the mass equation, and the defintion of α, and find: where M n = ξ 2 1 dθ dξ ) n 1 ) 3 n GM R = ) ξ 1 M n R n 0 [n + 1)K]n, 16) 4πG and R n = ξ 1. We note that this gives a mass-radius relation for a given polytropic index. For n=3, the mass is not a function of the radius it s only a function of K, the scale factor for the pressure-density relationship: M = 4πM 3 K πg) 3/2 17) Thus, for a given K there is only a single possible value of mass for an n = 3 polytrope. Another special case is the n = 1 polytrope, in which only one value of R can exist, R = R K 1/2 1 2πG) and it is independent of M. In between, R 3 n 1, 18) M n 1 meaning that more massive stars are denser. 3

Now, we can pull out K from the mass-radius relation, and use P c = kρ 1+ 1 n c, and put all the remaining n dependence into B n, which turns out always to be approximately 1, and we can get: yielding the result that P c M 2/3 ρ 4/3 c 2 The Chandrasekhar mass P c = 4π) 1/3 B n GM 2/3 ρ 4/3 c, 19) for nearly any equation of state. Despite the fact that white dwarfs are exotic stars, and inherently quantum mechanical, they actually provide one of the easiest challenges for stellar physics. This is because white dwarfs are the one class of stars with very simple equations of state. Let s first consider the special case where the white dwarf is low mass, so purely non-relativistic, and has a sufficiently low temperature that thermal motions can be ignored. It then follows an n = 1.5 polytrope, since its equation of state is exactly that for a non-relativistic degenerate electron gas. Then, R M 1/3, and barρ MR 3 M 2. Eventually, the density becomes so large that the electrons must become relativistic. Then, the solution will approach a n = 3 polytrope, for which only one mass is possible for a given value of K). We can then take the value of K for degenerate relativistic electrons from a few weeks/chapters back in the notes/book, and find: M ch = M 3 1.5 4π hc Gm 4/3 H ) 3/2 µ 2 e, 20) which yields M ch = 1.46M for matter with equal numbers of neutrons and protons. For iron, the number will be about 15% lower. For pure hydrogen, it would be a factor of 4 higher, but of course in a star made of 5M of hydrogen, fusion would take place, so the point is moot. 3 The Eddington Luminosity Is there some fundamental limit to the rate at which energy can be produced in a star, or the rate at which energy can be transported through a star, without a violation of the basic assumptions that go into what we ve done here? The answer is yes; although for the latter question, we can consider alternative means of energy transport. So, let s start with the equation for radiation pressure: P rad = 1 3 at 4. We can then differentiate this with respect to radius; substitute into equation 3; and divide by equation 1. This yields: rad = κf 4πcGm. This gives an upper limit for the stellar luminosity and for the star to be in hydrostatic equilibrium, since rad < 1 since the radiation pressure must everywhere be less than the total pressure or the gas pressure would be negative), 4

and hence cannot change by more than the total pressure). Then, κf < 4πGcm, and L < 4πGcm κ. Now, at high heat flux, or at low opacity, this condition may be violated, and then we lost hydrostatic equilibrium, at least locally. In the interior of a star, convection will occur, and in the exterior, a wind will be driven. We also get a local maximum rate of energy generation at the center of the star, q c < 4πcG κ. The maximum luminosity, called the Eddington luminosity can be re-written as: ) ) 1 M κ L EDD = 3.2 10 4 L 21) M κ es Since the solar luminosity is about 4 10 26 W, or 4 10 33 ergs/sec, the Eddington luminosity can also be written as 1.38 10 31 W or 1.38 10 38 ergs/sec. 4 Eddington s standard model This treatment of radiation pressure as a fundamental quantity can help motivate an approach to a key historical simplifcation of the equations of stellar physics, one also developed by Eddington. Eddington found that with the assumption that the total pressure is proportional to gas pressure everywhere in the star, then an n = 3 polytrope can be taken as the solution for stellar structure in all cases. This leads to mass being a function of K, with K a function of the ratio of radiation to gas pressure. Eddington justified the assumption that gas pressure is proportional to radiation pressure with an approach that is not rigorous, but which was instrumental in getting some sort of stellar model to work. He first wrote: F m = η L M, 22) defining η as the ratio between luminosity and mass produced within a radius and the same quantity at the outside of the star. It is clear that η increases inwards in the star. He next noted that rad, which we see above. Clearly, since the temperature is larger inwards and Kramer s opacity decreases sharply with increasing temperature, κ should increase outwards. Eddington then decided to calculate what would happen if one assumed κη κ s, where κ s is a constant. Then: = κf 4πcGm P rad = κ sl P, 23) 4πcGm where L = L EDD 1 β), where β = Pgas P, and we now have constant β. at We can then define 4 31 β) = R βµ ρt. 5

and, This yields three equations: [ ] 1/3 3R1 β T = ρ 1/3, 24) aµβ P = Kρ 4/3, 25) [ 3R 4 ] 1/3 1 β) K = aµ 4 β 4. 26) Since K is a constant, we get an n = 3 polytrope. Then, we get M = 4πM 3 K πg) 3/2, and we can substitute in M3 and R, and we get Eddington s quartic equation: 1 β = 3 10 3 M M ) 2 µ 4 β 4. 27) An interesting point is that only in a narrow range of M is β anything other than very close to 0 or very close to 1 i.e. only for a narrow range of M is the star anything other than purely gas pressure dominated or purely radiation pressure dominated and this range turns out to be the range for which stars actually exist. Two key general results can be found: 1) Massive stars are radiation pressure dominated to the extent that very massive stars come very close to the Eddington luminosity and probably drive stellar winds which they are observed to do). 2) If we take Eddington s quartic equation, and substitude L/L Edd for 1 β), then note that L EDD M, we find that L M 3, which is a pretty good approximation to reality, at least for massive main sequence stars. This was not known from observations at Eddington s time. It is fascinating to see how far Eddington came toward describing stellar structure reasonably accurately with essentially no understanding of the nuclear physics that provided the energy source! 5 The Cowling approximation point source models An alternative simplification one which can better describe the lower main sequence is that all of the nuclear power is generated from a point in the center of the star. Then, we can set F = L, and let F be a constant. The solutions are a bit more complicated and require numerical integration, but they were numerical integrations 2 that could be done in the 1930 s. 2 Numerical integration basically consists of plotting a function, then fitting rectangles or trapezoids, or other shapes along the curve, and adding up their areas. Some methods use some tricks to reduce the number of shapes that need to be drawn, but increase the complication of the shape so that it s a closer approximation of the curve, but the basic underlying idea is the same. 6

Next, we set κ = κ 0 ρ a T b, to define the opacity, since it need not be related to the radiation pressure. Then: dr = Gmρ r 2 28) rad dr = κ 0L ρ a+1 T b c 4πr 2 29) dm dr = 4πr2 ρ. 30) This can be integrated numerically for any given opacity model. For the simplest opacity model, a = b = 0, constant opacity relevant, e.g. in cases where electron scattering dominates), some additional progress can be made in solving the equations in a more straightforward way. We won t go into this level of detail in this course by we note that the resulting solutions give a steeper M L relationship than the Eddington model and thus they agree better with real low mass stars. This is not surprising the Eddington model implicitly assumes that fusion is taking place all the way out in the envelope of the star, whereas low mass stars barely have high enough density and temperature to have any fusion at all putting all the fusion in their cores is a better approximation. 7