chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · chapter 5 nuclear reactions in...

32
Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W. What follows is a brief outline as given in the lecture presentation. For a star in thermal equilibrium, an internal energy source is required to balance the energy loss from the surface in the form of radiation. This energy source is provided by nuclear reactions that take place in the deep interior, where the temperature and density are suciently high. Apart from energy generation, another important eect of nuclear reactions is that they change the composition by transmutations of one element into another. 5.1 Nuclear reactions and composition changes Consider a nuclear reaction whereby two nuclei (denoted I and J) combine to produce two other nuclei (K and L): I + J K + L (5.1) Many, though not all, reactions are of this type, and the general principles discussed here also apply to reactions involving dierent numbers of nuclei. Each nucleus is characterized by two integers, the baryon number or mass number A i and the charge Z i . Nuclear charges and baryon numbers must be conserved during a reaction, i.e. for the example above: Z i + Z j = Z k + Z l and A i + A j = A k + A l (5.2) Some reactions involve electrons or positrons, or neutrinos or antineutrinos. In that case the lepton number must also be conserved during the reaction. Therefore any three of the reactants uniquely determine the fourth. The reaction rate – the number of reactions taking place per cm 3 and per second – can therefore be identified by three indices. We denote the reaction rate of reaction (5.1) as r ij ,k . The rate at which the number density n i of nuclei I changes with time can generally be the result of dierent nuclear reactions, some of which consume I as above, and others that proceed in the reverse direction and thereby produce I . If we denote the rate of reactions of the latter type as r kl,i , we can write dn i dt = j ,k r ij ,k + k,l r kl,i (5.3) 53

Upload: dangthien

Post on 14-Apr-2018

217 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Chapter 5

Nuclear reactions in stars

The contents of this chapter are covered by Chapter 18 ofK &W. What follows is a brief outline as given in the lecture presentation.

For a star in thermal equilibrium, an internal energy sourceis required to balance the energy lossfrom the surface in the form of radiation. This energy sourceis provided bynuclear reactionsthattake place in the deep interior, where the temperature and density are sufficiently high. Apart fromenergy generation, another important effect of nuclear reactions is that they change the compositionby transmutations of one element into another.

5.1 Nuclear reactions and composition changes

Consider a nuclear reaction whereby two nuclei (denotedI and J) combine to produce two othernuclei (K andL):

I + J→ K + L (5.1)

Many, though not all, reactions are of this type, and the general principles discussed here also applyto reactions involving different numbers of nuclei. Each nucleus is characterized by two integers, thebaryon number or mass numberAi and the chargeZi. Nuclear charges and baryon numbers must beconserved during a reaction, i.e. for the example above:

Zi + Z j = Zk + Zl and Ai + A j = Ak + Al (5.2)

Some reactions involve electrons or positrons, or neutrinos or antineutrinos. In that case the leptonnumber must also be conserved during the reaction. Therefore any three of the reactants uniquelydetermine the fourth. The reaction rate – the number of reactions taking place per cm3 and per second– can therefore be identified by three indices. We denote the reaction rate of reaction (5.1) asr i j,k.

The rate at which the number densityni of nuclei I changes with time can generally be the resultof different nuclear reactions, some of which consumeI as above, and others that proceed in thereverse direction and thereby produceI . If we denote the rate of reactions of the latter type asrkl,i , wecan write

dni

dt= −∑

j,k r i j,k +∑

k,l rkl,i (5.3)

53

Page 2: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

The number density of nucleusI is related to the mass fractionXi by ni = Xi ρ/(Aimu), so that we canwrite the rate of change of the mass fraction due to nuclear reactions as

dXi

dt= Ai

mu

ρ

( −∑

j,k r i j,k +∑

k,l rkl,i)

(5.4)

For each nuclear speciesi one can write such an equation, describing the composition change ata particular mass shell inside the star (with densityρ and temperatureT) resulting from nuclearreactions. In the presence of internal mixing (in particular of convection, Sec. 4.4.3) the redistributionof composition between different mass shells should also be taken into account.

5.2 Nuclear energy generation

• energy released per reaction, e.g.I + J→ K + L

Qi j,k = (mi +mj −mk −ml)c2 (, 0 becausemi , Aimu)

• energy generation rate (erg g−1 s−1) from this reaction: ǫi j,k =Qi j,kr i j,k

ρ⇒ total nuclear

energy gerenation rate:

ǫnuc =∑

i, j,k

ǫi j,k =∑

i, j,k

Qi j,k

ρr i j,k (5.5)

• Q corresponds to difference in binding energy of the nuclei involved (usually expressed inMeV)

EB(AZ) := [(A− Z)mn + Z ·mp −m(AZ)] · c2

Figure 5.1. Binding energy per nucleonEB/A

54

Page 3: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

m(AZ) = mass of nucleus with massA and proton numberZ

∆m= (A− Z)mn + Zmp − M(AZ) = “mass defect”

5.3 Thermonuclear reaction rates

• nuclear reaction rate (no. of reactions s−1 cm−3) at relative velocityv:

r i j (v) :=1

1+ δi jni n j vσ(v)

• in stellar gas: Maxwell-Boltzmann velocity distributionφ(v)

φ(v) = 4πv2( m2πkT

)3/2exp

(

−mv2

2kT

)

r i j =1

1+ δi jnin j〈σv〉 = 1

1+ δi jnin j

∫ ∞

0φ(v)vσ(v)dv (5.6)

• hence〈σv〉 = f (T) only depends on the temperature:

〈σv〉 = (8/πm)12 (kT)−

32

∫ ∞

0σ(E)E exp

(

− EkT

)

dE (5.7)

5.3.1 Coulomb barrier

Figure 5.2. Depiction of the combined nuclear and Coulomb potential.

Coulomb potential

EC =Z1Z2e2

r⇒ typical Coulomb barrierEC(rn) ≈ Z1Z2 MeV

Compare to typical particle energy (107 K): 〈Ekin〉 = 32kT ≈ 1 keV

55

Page 4: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.3.2 Tunnel effect

• even ifE≪ EC(rn)⇒ finite probability that projectile penetrates repulsive Coulomb barrier

• tunneling probability: P = exp(−b · E−1/2)

approximately P ≈ exp(−rc/λ)

rc =Z1Z2e2

E

λ =~

p=

~√

2mE(de Broglie wavelength)

• quantitatively: P = exp(−2πη)

2πη = 2πZ1Z2e2(1

2m)1/2

~E1/2= 31.29 · Z1Z2

(AE

)

12

(E in keV)

A =A1A2

A1 + A2(reduced mass inmu)

5.3.3 Cross sections

• classically: σ = π(R1 + R2)2

• QM: σ = πλ2 with λ =~

p∝ 1√

E

Figure 5.3. Example of the dependence of the reaction cross section withenergyE for the3He+4He→ 7Be+γreaction. Althoughσ varies very strongly with energy, and becomes immeasurableat very lowE, theS(E)-factor is only a very weak function ofE and (in the case of this reaction) can be safely extrapolatedto the lowenergies that are relevant for nuclear reactions in stars (red vertical bar).

56

Page 5: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

0 20 40 60 80 10010−30

10−20

10−10

10+0

E (keV)

E (keV)

E (keV)

12C + 1H −> 13N + γ

T = 3.0 x 107 Kexp(−E/kT)

exp(−b/E1/2)

0 20 40 60 80 1000.0

0.2

0.4

0.6

0.8

1.0

10−20

E (keV)

E (keV)

E (keV)

12C + 1H −> 13N + γT = 3.0 x 107 K

Figure 5.4. Example of the Gamow peak for the12C + p → 13N + γ reaction. The left panel shows asdasked lines the tunnel probablity factor (exp(−b/E1/2)) and the tail of the Maxwell distribution (exp(−E/kT))for three values of temperature:T = 2.0 × 107 K (lower, black), 2.5 × 107 K (middle, red) and 3.0 × 107 K(upper, blue). The solid lines are the resulting Gamow peaks. To appreciate the sharpness of the peak, andthe enormous sensitivity to temperature, the same curves are plotted on a linear scale in the right panel. TheGamow peak forT = 2.0× 107 K is barely visible at∼ 30 keV, while that for 3.0× 107 K is off the scale.

• tunneling probability:P = exp(−b/√

E)

DefineS(E) as follows:

σ(E) := S(E)exp(−b/

√E)

E(5.8)

S(E) = “astrophysicalS-factor”

• slowly varying function ofE (unlikeσ !)

• intrinsic nuclear properties (e.g. resonances)

• extrapolated from measurements at highE

5.3.4 The Gamow peak

Combining eqs. (5.7) and (5.8)⇒

〈σv〉 = (8/πm)12 (kT)−

32

∫ ∞

0S(E) exp

(

− EkT− b√

E

)

dE (5.9)

≈ (8/πm)12 (kT)−

32 S(E0)

∫ ∞

0exp

(

− EkT− b√

E

)

dE (5.10)

where the integrand, with two competing exponentials, describes the so-calledGamow peak(seefigure below). See K&W chapter 18 for a discussion of the properties.

For different reactions,b ∝ Z1Z2

A1A2A1+A2

, so that a reaction between heavier nuclei (largerA andZ) will have a much lower rate at constant temperature. In other words, it will need a higherT tooccur at an appreciable rate.

57

Page 6: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.3.5 Temperature dependence of reaction rates

Approximate Gamow peak by Gaussian:

height= exp(−3E0

kT) ∝ exp(− b

23

T13

)

width = 4

E0kT3∝ b

13 T

56

So that from eq. (5.10):

〈σv〉 ∝ b13 T−

23 exp(− b

23

T13

) (5.11)

This means that

• 〈σv〉 drops strongly with increasing Coulomb barrier

• 〈σv〉 increasesvery strongly with temperature

Approximate〈σv〉 ≈ 〈σv〉0 Tν, where ν :=d log〈σv〉d logT

≫ 1

example (T = 1.5 · 107K):

reaction 〈σv〉 ∝ EC

p+ p T3.9 0.55 MeVp+ 14N T20 2.27 MeVα + 12C T42 3.43 MeV16O+ 16O T182 14.07 MeV

Sensitivity of reaction rates toT and toZ1Z2 implies:

• nuclear burning of H, He, C, etc are well separated in temperature

• only few reactions occur at same time

Energy generation rate

combining expressions forǫi j = Qi jr i j

ρandr i j = ni n j〈σv〉 ⇒

energy generation rate (for two-particle reactions):

ǫi j =Qi j

m2u

XiX j

AiA jρ 〈σv〉 ≈ ǫ0 Xi X j ρTν (5.12)

whereǫ0 depends strongly on the Coulomb barrier (ZiZ j)

58

Page 7: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.4 The main nuclear burning cycles

5.4.1 Hydrogen burning

Net effect of hydrogen burning:

41H→ 4He+ 2 e+ + 2ν (Q = 26.73 MeV)

energy release per gram:Q/m(4He)= 6.4× 1018 erg/g

• two neutrinos released per4He (weak interactions, p→ n)⇒ decrease of the effectiveQ value

• four-particle reaction extremely unlikely⇒ chain of reactions necessary

• first reaction is the pp reaction:

p+ p→ D + e+ + ν

involves simultaneousβ-decay of p⇒ extremely small cross section, 10−20× that of stronginteraction. The reaction rate is unmeasurably small⇒ only known from theory

5.4.2 The p-p chains

After some D is produced, several reaction paths are possible:

1H + 1H→ 2H + e+ + ν2H + 1H→ 3He+ γ

��

��

���

HH

HH

HHj

3He+ 3He→ 4He+ 2 1H

ppI

3He+ 4He→ 7Be+ γ

��

��

���

HH

HH

HHj

7Be+ e− → 7Li + ν7Li + 1H→ 4He+ 4He

ppII

7Be+ 1H→ 8B + γ8B→ 8Be∗ + e+ + ν8Be∗ → 4He+ 4He

ppIII

For T > 8× 106 K ⇒ reactions in equilibrium⇒ overall reaction rate= rpp (slowest reaction)

Energy production by pp chain

Neutrino release:ppI: 0.53 MeV per4HeppII: 0.81 MeV per4HeppIII: 6.71 MeV per4He⇒ effectiveQ value depends on temperature

59

Page 8: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.4.3 The CNO cycle

• net effect: 41H→ 4He+ 2 e+ + 2ν

• at high enoughT: reactions in equilibrium

• total number of CNO-nuclei is conserved

• define lifetime of nuclear speciesi against reacting with particlej:

τ j(i) :=

ni

[dni/dt] j

=1

n j〈σv〉i j(5.13)

atT ≃ 2 · 107 K: τp(15N) ≪ τp(13C) < τp(12C)≪ τp(14N)≪ τstar

i.e. 35 yr 1600 yr 6600 yr 9× 105 yr

• at high enoughT, 12C,13C,14N and15N will seek equilibrium abundances, given bydn(12C)/dt =dn(13C)/dt = ..., etc⇒

(

n(12C)n(13C)

)

eq=〈σv〉13

〈σv〉12=τp(12C)

τp(13C), etc (5.14)

• overall rate of CNO-cycle deteremined by the14N(p, γ)15O reaction

• apart from4He, 14N is the major product of the CNO-cycle

CNO cycle vs pp chain

See K&W figure 18.8.

Temperature sensitivity approximately:

ǫpp ∝ X2 ρT4 (5.15)

ǫCNO ∝ X X14ρT18 (5.16)

60

Page 9: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.4.4 Helium burning

No stable nucleus withA = 8⇒ helium burning occurs in two steps:

1. α + α↔ 8Be builds up small equilibrium concentration of8Be

2. 8Be+ α→ 12C∗ → 12C+ γ becomes possible atT ≈ 108 K due to a resonance in12C

(predicted by Fred Hoyle in 1954 on the basis of the existenceof carbon in the Universe)

Net effect is calledtriple-α reaction:

3α→ 12C+ γ (Q = 7.275 MeV)

energy release per gram:Q/m(12C) = 5.9× 1017 erg/g (≈ 1/10 of H-burning)

When sufficient12C has been created, also:

12C+ α→ 16O+ γ (Q = 7.162 MeV)

reaction rate of12C(α, γ)16O is uncertain⇒ affects final C/O ratio...

Approximately, atT ≈ 108 K:

ǫ3α ∝ Y3 ρ2 T40 (!)

ǫαC ∝ Y X12ρT20

12C

16O

0.00.20.40.60.81.00.0

0.2

0.4

0.6

0.8

1.0

X(4He)

mas

s fr

actio

n

Figure 5.5. Dependence of the mass fractions of12C and16O on4He during He-burning, for typical conditionsin intermediate-mass stars.

61

Page 10: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5.4.5 Carbon burning and beyond

The following nuclear burning processes occur at successively higher temperatures in the cores ofmassive stars:

Carbon burning occurs atT9(≡ T/109 K) ∼> 0.5 mainly by the following reaction channels:

12C+ 12C→ 24Mg∗ → 20Ne+ α + 4.6 MeV (∼ 50%)→ 23Na+ p + 2.2 MeV (∼ 50%)

Many side reactions occur with the released p andα particles, e.g.23Na(p, α)20Ne,20Ne(α, γ)24Mg,and many others. The composition after carbon burning is mostly 16O, 20Ne and24Mg (together95% by mass fraction). These most abundant nuclei have equalnumbers of protons and neutron,but some of the side reactions produce neutron-rich isotopes like 21,22Ne,23Na and25,26Mg, sothat after C burning the overall composition has a ‘neutron excess’ (n/p > 1, orµe > 2).

Neon burning is the next cycle, occurring at lower temperature (T9 ≈ 1.5) than oxygen burningbecause16O is a very stable ‘double-magic’ nucleus (N = Z = 8). It occurs by a combinationof photodisintegration andα-capture reactions:

20Ne+ γ↔ 16O+ α − 4.7 MeV20Ne+ α→ 24Mg + γ + 9.3 MeV

The first reaction is endothermic, but effectively the two reactions combine to 220Ne→ 16O+24Mg with a net energy generation> 0. The composition after neon burning is mostly16O and24Mg (together 95% by mass).

Oxygen burning occurs atT9 ≈ 2.0 mainly by the reaction channels:

16O+ 16O→ 32S∗ → 28Si+ α + 9.6 MeV (∼ 60%)

31P+ p + 7.7 MeV (∼ 40%)

together with many side reactions with the released p andα particles, as was the case for C-burning. The composition after oxygen burning is mostly28Si and32S (together 90% by mass)

Silicon burning does not occur by28Si + 28Si, but instead by a series of photodisintegration (γ, α)andα-capture (α, γ) reactions whenT9 ∼> 3. Part of the silicon ‘melts’ into lighter nuclei, whileanother part captures the released4He to make heavier nuclei:

28Si (γ, α) 24Mg (γ, α) 20Ne (γ, α) 16O (γ, α) 12C (γ, α) 2α28Si (α, γ) 32S (α, γ) 36Ar (α, γ) 40Ca (α, γ) 44Ti (α, γ) . . . 56Ni

Most of these reactions are in equilibrium with each other, and their abundances can be de-scribed by nuclear equivalents of the Saha equation for ionization equilibrium. ForT > 4·109 Ka state close tonuclear statistical equilibrium (NSE)can be reached, where the most abundantnuclei are those with the lowest binding energy, constrained by the total number of neutrons andprotons present. The final composition is then mostly56Fe because p/n < 1 (due toβ-decaysand e−-captures).

62

Page 11: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Chapter 6

Stellar models and stellar stability

In the previous chapters we have reviewed the most importantphysical processes taking place instellar interiors, and we derived the differential equations that determine the structure and evolution ofa star. By putting these ingredients together we can construct models of spherically symmetric stars.Because the complete set of equations is highly non-linear and time-dependent, their full solutionrequires a complicated numerical procedure. This is what isdone in detailed stellar evolution codes,the results of which will be described in later chapters. We will skip any details about the numericalmethods commonly used in such codes – for those interested, some of those details may be found inChapter 11 of K&W.

The main purpose of this chapter is to briefly analyse the differential equations of stellar evolutionand their boundary conditions, and to see how the full set of equations can be simplified in some casesto allow simple or approximate solutions - so-calledsimple stellar models. Finally we address thequestion of the stability of stars – whether the solutions tothe equations yield a stable structure ornot.

6.1 The differential equations of stellar evolution

Let us collect and summarize the differential equations for stellar structure and evolution that we havederived in the previous chapters, i.e. eqs. (2.5), (2.10), (2.36), (4.5) and (5.4):

∂r∂m=

14πr2ρ

(6.1)

∂P∂m= − Gm

4πr4− 1

4πr2

∂2r∂t

(6.2)

∂l∂m= ǫnuc− ǫν − T

∂s∂t

(6.3)

∂T∂m= − Gm

4πr4

TP∇ with ∇ =

∇rad =3κ

16πacGlP

mT4if ∇rad ≤ ∇ad

∇ad+ ∆∇ if ∇rad > ∇ad

(6.4)

∂Xi

∂t=

Aimu

ρ

(

−∑

j,k

r i j,k +∑

k,l

rkl,i)

[+ mixing terms] i = 1 . . .N (6.5)

Note that eq. (6.2) is written in its general form, without pre-supposing hydrostatic equilibrium. Ineq. (6.3) we have replaced∂u/∂t−(P/ρ2)∂ρ/∂t by T∂s/∂t, according to the combined first and second

63

Page 12: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

laws of thermodynamics. Eq. (6.4) is generalized to includeboth the cases of radiative and convectiveenergy transport. The term∆∇ is the superadiabaticity of the temperature gradient that must followfrom a theory of convection (in practice, the mixing length theory); for the interior one can take∆∇ = 0 except in the outermost layers of a star. Finally, eq. (6.5)has been modified to add ‘mixingterms’ that describe the redistribution (homogenization)of composition in convective regions. ThereareN such equations, one for each nucleus (isotope) indicated bysubscripti.

The set of equations above comprise 4+ N partial differential equations that should be solvedsimultaneously. Let us count the number of unknown variables. Making use of the physics discussedin previous chapters, the functionsP, s, κ, ∇ad, ∆∇, ǫnuc, ǫν and the reaction ratesr i j,k can all beexpressed as functions ofρ, T and compositionXi. We are therefore left with 4+N unknown variables(r, ρ, T, l and theXi) so that we have a solvable system of equations.

The variablesr, ρ, T, l and Xi appearing in the equations are all functions of twoindependentvariables,m and t. We must therefore find a solution to the above set of equations on the interval0 ≤ m≤ M for t > t0, assuming the evolution starts at timet0. (Note thatM generally also depends ont in the presence of mass loss.) A solution therefore also requires specification ofboundary conditions(atm= 0 andm= M) and ofinitial conditions, for exampleXi(m, t0).

6.1.1 Timescales and initial conditions

Let’s further analyse the equations. Three kinds of time derivatives appear:

• ∂2r/∂t2 in eq. (6.2), which describes hydrodynamical changes to thestellar structure. Theseoccur on the dynamical timescaleτdyn which as we have seen is very short. Thus we cannormally assume hydrostatic equilibrium and∂2r/∂t2 = 0, in which case eq. (6.2) reduces tothe ordinary differential equation (2.12). Note that HE was explicitly assumed in eq. (6.4).

• T∂s/∂t in eq. (6.3), which is often written as an additional energy generation term,ǫgr:

ǫgr = −T∂s∂t= −∂u∂t+

P

ρ2

∂ρ

∂t

It describes changes to the thermal structure of the star, which can result from contraction(ǫgr > 0) or expansion (ǫgr < 0) of the layers under consideration. Such changes occur on thethermal timescaleτKH. If a star evolves on a much slower timescale thanτKH then ǫgr ≈ 0and the star is in thermal equilibrium. Then also eq. (6.3) reduces to an ordinary differentialequation, eq. (2.37).

• ∂Xi/∂t in eqs. (6.5), describing changes in the composition. For the most abundant elements –the ones that affect the stellar structure – such changes normally occur on the longest, nucleartimescaleτnuc.

Because normallyτnuc ≫ τKH ≫ τdyn, composition changes are usually very slow compared to theother time derivatives. In that case eqs. (6.5) decouple from the other four equations (6.1–6.4), whichcan be seen to describe thestellar structurefor a given compositionXi(m).

For a star in both HE and TE (also called ‘complete equilibrium’), the stellar structure equa-tions (6.1–6.4) become a set of ordinary differential equations, independent of time. In that case it issufficient to specify the initial composition profilesXi(m, t0) as initial conditions. This is the case forso-calledzero-age main sequencestars: the structure at the start of the main sequence depends onlyon the initial composition, and is independent of the uncertain details of the star formation process, avery fortunate circumstance!

64

Page 13: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

If a star starts out in HE, but not in TE, then the time derivative represented byǫgr remains in theset of structure equations. One would then also have to specify s(m, t0) as an initial condition. This isthe case if one considers pre-main sequence stars. Fortunately, as we shall see later, in this case thereis also a simplifying circumstance: pre-main sequence stars start out as fully convective gas spheres.This means that their temperature and pressure stratification is nearly adiabatic, so thatscan be takenas constant throughout the star. It then suffices to specify the initial entropy.

6.2 Boundary conditions

The contents of this section are covered by Chapter 10 ofK &W.

6.2.1 Central boundary conditions

At the centre (m= 0), both the density and the energy generation rate must remain finite. Therefore,bothr andl must vanish in the centre:

m= 0 ⇒ r = 0, l = 0.

However, nothing is known a priori about the central values of P andT. Therefore the remaining twoboundary conditions must be specified at the surface rather than the centre.

The behaviour of e.g.P andT close to the centre can be described by a Taylor expansion, e.g.

P = Pc +m

[

dPdm

]

c+ 1

2m2[

d2P

dm2

]

c+ . . .

(see K&W sec. 10.1)

6.2.2 Surface boundary conditions

At the surface (m= M, or r = R), the boundary conditions are generally much more complicated thanat the centre. A few options are, at increasing level of sophistication.

• The simplest option is to takeT = 0 andP = 0 at the surface (the ‘zero’ boundary conditions).However, in realityT andP never become zero because the star is surrounded by an interstellarmedium with very low but finite density and temperature.

• A better option is to identify the surface with thephotosphere, which is where the bulk of theradiation escapes and which corresponds with the visible surface of the star. The photosphericboundary conditions approximate the photosphere with a single surface at optical depthτ = 2

3.We can write

τ(R) =∫ ∞

Rκρdr ≈ κph

∫ ∞

Rρdr, (6.6)

whereκph is an average value of opacity over the atmosphere (the layers above the photosphere).In the atmosphere we also have

dPdr= −GM

R2ρ ⇒ P(R) ≈ GM

R2

∫ ∞

Rρdr. (6.7)

65

Page 14: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Sinceτ(R) = 23 andT(R) ≈ Teff we can combine the above equations to write the photospheric

boundary conditions as:

P =23

GM

κphR2and L = 4πR2σT4 (6.8)

• The problem with the photospheric boundary conditions above is that the radiative diffusionapproximation breaks down whenτ ∼< a few. The best solution is therefore to fit adetailedstellar atmosphere modelto an interior shell (atτ > 2

3) below which radiative diffusion is stillvalid. This is a more complicated and time-consuming approach, and in many (but not all)practical situations the photospheric conditions are sufficient.

6.2.3 Effect of surface boundary conditions on stellar structure

Consider a radiative outer envelope, containing only a small fraction of the mass and no energysources. In that casel = L andm≈ M, so that

dTdP=

TP∇rad = const· L

T3

Assume a simple opacity law:κ = κ0PaTb (e.g. Kramers:a = 1, b = −4.5; or electron scattering:a = b = 0). We can then integratedT/dP to give

T4−b = B (P1+a +C)

whereB ∝ L/M = constant andC is an integration constant, determined by the boundary conditions.E.g. for the Kramers opacity:T8.5 = B (P2+C), so that all solutions approachT ∝ P0.235 for P≫

√C

Read K&W section 10.3, and compare their Fig. 10.2 to Fig. 6.1below.

radiative envelope corresponds to solutions withC > 0

• requiresTeff larger than critical value (≈ 9000 K)

• the surface BC only has a small effect on the envelope structure, so that assuming thephotospheric BC’s is okay

convective envelopecorresponds to solutions withC < 0

• occurs forTeff ∼< 9000 K

• the envelope structure is sensitive to surface BC: a small change inTeff has a large effecton the depth of the convective envelope!

• for small enoughTeff the whole star can become convective (leading to the Hayashilinein the HRD, see Sect. 8.1.1)

66

Page 15: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

5 10 15

4

5

6

7

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

log P (dyn/cm2)

log

T (

K)

ZAMS models, Z = 0.02

Figure 6.1. Structure of detailed stellar models onthe zero-age main sequence in the logP, logT dia-gram. Each curve is for a different mass, from topto bottom: 16M⊙, 8M⊙, 4M⊙, 2M⊙, 1M⊙, 0.5M⊙and 0.25M⊙. The+ symbols on each curve indicate,for increasingP, the part of the envelope containing0.01M, 0.1M and 0.5M (so most of theT andP vari-ation in the envelope occurs in the outer 1 % of themass). The dotted (green) parts of each curve indi-cate radiative regions of the star, the solid (red) partsindicate convective regions.

Stars withM ≤ 1 M⊙ have lowTeff and thereforeconvective envelopes. The depth of the convectiveenvelope increases strongly with decreasing surfacetemperature (and thus with decreasing mass); the0.25M⊙ star is completely convective. On the otherhand, more massive stars have higherTeff and mostlyradiative envelopes, except for small convective lay-ers near the surface caused by partial ionization.

6.3 Equilibrium stellar models

For a star in both hydrostatic and thermal equilibrium, the four partial differential equations for stellarstructure (eqs. 6.1–6.4) reduce to ordinary, time-independent differential equations. We can furthersimplify the situation somewhat, by ignoring possible neutrino losses (ǫν) which are only important invery late stages of evolution, and ignoring the superadiabaticity of the temperature gradient in surfaceconvection zones. We then arrive at the following set of structure equations that determine the stellarstructure for a given composition profileXi(m):

drdm=

14πr2ρ

(6.9)

dPdm= − Gm

4πr4(6.10)

dldm= ǫnuc (6.11)

dTdm= − Gm

4πr4

TP∇ with ∇ =

∇rad =3κ

16πacGlP

mT4if ∇rad ≤ ∇ad

∇ad if ∇rad > ∇ad

(6.12)

We note that the first two equations (6.9 and 6.10) describe the mechanical structureof the star, andthe last two equations (6.11 and 6.12) describe thethermal and energeticstructure. They are coupledto each other through the fact that, for a general equation ofstate,P is a function of bothρ andT.

Although simpler than the full set of evolution equations, this set still has no simple, analyticsolutions. The reasons are that, first of all, the equations are very non-linear: e.g.ǫnuc ∝ ρTν withν ≫ 1, andκ is a complicated function ofρ andT. Secondly, the four differential equations arecoupled and have to be solved simultaneously. Finally, the equations have boundary conditions atboth ends, and thus require iteration to obtain a solution.

67

Page 16: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

It is, however, possible to make additional simplifying assumptions so that under certain circum-stances an analytic solution or a much simpler numerical solution is possible. We shall discuss twosuch simplifying approaches. In the first approach, we assume a specific,polytropic relation betweenpressure and density,

P = K ργ.

Since in this caseP does not depend onT, the mechanical structure of a stellar model can be computedin a simple way, independent of its thermal and energetic properties, by solving eqs. (6.9) and (6.10).Such models are calledpolytropic models.

In the second approach we consider simple scaling relationsbetween stellar models with differentmasses and radii, but all having the same (or a very similar) relative density distributions. If a detailednumerical solution can be computed for one particular star,these so-calledhomology relationscan beused to find an approximate model for another star.

6.4 Polytropic models

The contents of this section are covered by Chapter 19 ofK& W,sections 19.1–7 and 19.9, and by the second part of the computer practicum.

6.5 Homology relations

Solving the stellar structure equations almost always requires heavy numerical calculations, suchas applied in detailed stellar evolution codes. However, there is often a kind of similarity betweenthe numerical solutions for different stars. These can be approximated by simple analyticalscalingrelations known ashomology relations. In past chapters we have already applied simple scalingrelations based on rough estimates of quantities appearingin the stellar structure equations. In thissection we will put these relations on a firmer mathematical footing.

The requirements for the validity of homology are very restrictive, and hardly ever apply to re-alistic stellar models. However, homology relations can offer a rough but sometimes very helpfulbasis for interpreting the detailed numerical solutions. This applies to models for stars on themainsequenceand to so-calledhomologous contraction.

Definition Compare two stellar models, with massesM1 and M2 and radiiR1 andR2. All interiorquantities in star 1 are denoted by subscript ’1’ (e.g. the mass coordinatem1), etc. Now considerso-calledhomologous mass shellswhich have the same relative mass coordinate,x ≡ m/M, i.e.

x =m1

M1=

m2

M2(6.13)

The two stellar models are said to behomologousif homologous mass shells within them arelocated at the same relative radiir/R, i.e.

r1(x)R1=

r2(x)R2

orr1(x)r2(x)

=R1

R2(6.14)

for all x.

68

Page 17: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Comparing two homologous stars, the ratio of radiir1/r2 for homologous mass shells is constant. Inother words, two homologous stars havethe same relative mass distribution, and therefore (as weshall prove shortly) the same relative density distribution.

All models have to obey the stellar structure equations, so that the transition for one homologousmodel to another has consequences for all other variables. We start by analysing the first two structureequations.

• The first stellar structure equation (6.9) can be written forstar 1 as

dr1

dx=

M1

4πr12ρ1

(6.15)

If the stars are homologous, then from eq. (6.14) we can substitute r1 = r2 (R1/R2) and obtain

dr2

dx=

M2

4πr22ρ2·[

ρ2

ρ1

M1

M2

(R2

R1

)3]

. (6.16)

We recognize the structure equation for the radius of star 2 (i.e. eq. 6.15 with subscript ‘1’ re-placed by ‘2’), multiplied by the factor in square brackets on the right-hand side. This equationmust hold generally, which is only the case if the factor in square brackets is equal to one forall values ofx, that is if

ρ2(x)ρ1(x)

=M2

M1

(R2

R1

)−3

. (6.17)

This must hold at any homologous mass shell, and therefore also at the centre of each star. Thefactor M R−3 is proportional to the average density ¯ρ, so that the density at any homologousshell scales with the central density, or with the average density:

ρ(x) ∝ ρc ∝ ρ (6.18)

Note that, therefore, any two polytropic models with the same indexn are homologous to eachother.

• We can apply a similar analysis to the second structure equation (6.10) for hydrostatic equilib-rium. For star 1 we have

dP1

dx= −GM1

2x

4πr14

(6.19)

so that after substitutingr1 = r2 (R1/R2) we obtain

dP1

dx= −GM2

2x

4πr24·[

(M1

M2

)2 (R2

R1

)4]

=dP2

dx·[

(M1

M2

)2 (R2

R1

)4]

, (6.20)

where the second equality follows because star 2 must also obey the hydrostatic equilibriumequation. Hence we havedP1/dx = C dP2/dx, with C equal to the (constant) factor in squarebrackets. Integrating we obtainP1(x) = C P2(x) + B, where the integration constantB = 0because at the surface,x→ 1, for both starsP→ 0. Thus we obtain

P2(x)P1(x)

=

(

M2

M1

)2 (

R2

R1

)−4

, (6.21)

69

Page 18: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

at any homologous mass shell. Again this must include the centre, so that for allx:

P(x) ∝ Pc ∝M2

R4. (6.22)

The pressure required for hydrostatic equilibrium therefore scales withM2/R4 at any homol-ogous shell. (Note that we found the same scaling of thecentralpressure withM andR fromour rough estimate in Sect. 2.2.)

We can combine eqs. (6.17) and (6.21) to show that two homologous stars must obey the followingrelation between pressure and density at homologous points,

P2(x)P1(x)

=

(

M2

M1

)2/3 (

ρ2(x)ρ1(x)

)4/3

, (6.23)

or

P(x) ∝ M2/3ρ(x)4/3. (6.24)

6.5.1 Main sequence homology

In order to obtain simple homology relations from the other structure equations, (6.11) and (6.12), wemust make additional assumptions.

• First, let us assume theideal gasequation of state,

P =RµρT.

Let us further assume that in each star thecomposition is homogeneous, so thatµ is a constantfor both stars, though not necessarily the same. We can then combine eqs. (6.17) and (6.21) toobtain a relation between the temperatures at homologous mass shells,

T2(x)T1(x)

=µ2

µ1

M2

M1

(

R2

R1

)−1

or T(x) ∝ Tc ∝ µMR

(6.25)

• Second, we will assume the stars are inradiative equilibrium. We can then write eq. (6.12) as

d(T4)dx

= − 3M

16π2ac

κl

r4(6.26)

This contains two as yet unknown functions ofx on the right-hand side,κ andl. We must there-fore make additional assumptions about the opacity, which we can very roughly approximateby a power law,

κ = κ0 ρaTb. (6.27)

For a Kramers opacity law, we would havea = 1 andb = −3.5. However, for simplicity let usassume aconstant opacitythroughout each star (but likeµ, not necessarily the same for both

70

Page 19: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

stars). Then a similar reasoning as was held above for the pressure, allows us to transformeq. (6.26) into an expression for the ratio of luminosities at homologous points,

(

T2(x)T1(x)

)4

=l2(x)l1(x)

M2

M1

κ2

κ1

(

R2

R1

)−4

⇒ l2(x)l1(x)

=

(

µ2

µ1

)4 (

M2

M1

)3 (

κ2

κ1

)−1

(6.28)

making use of eq. (6.25) to obtain the second expression. This relation also holds for the surfacelayer, i.e. for the total stellar luminosityL. Hencel(x) ∝ L and

L ∝ 1κµ4M3 (6.29)

This relation represents amass-luminosity relationfor a radiative, homogeneous star with con-stant opacity and ideal-gas pressure.

Note that we obtained this relation without making any assumption about the mode of energygeneration (and indeed, without even having to assume thermal equilibrium, because we havenot yet made use of eq. 6.11). We can thus expect a mass-luminosity relation to hold not onlyon the main-sequence, but for any star in radiative equilibrium. What this relation tells us isthat the luminosity depends mainly on how efficiently energy can be transported by radiation:a higher opacity gives rise to a smaller luminosity, becausethe intransparent layers work like ablanket wrapped around the star. In practice, for a star in thermal equilibrium (e.g. on the mainsequence) the power generated by nuclear reactionsLnuc adapts itself to the surface luminosityL, and thereby also the central temperature needed to make thenuclear reactions proceed at therate dictated byL.

Note also that the simple mass-luminosity relation (6.29) depends on the assumption of constantopacity. If we assume a Kramers opacity law, the mass-luminosity also depends (weakly) onthe radius. It is left as an exercise to show that, in this case

L ∝ µ7.5M5.5

R0.5.

This means that if the opacity is not a constant, there is a weak dependence of the luminosityon the mode of energy generation, through the radius dependence (see below).

• Finally, we can obtain a mass-radius relation by making use of eq. (6.11). We then have toassume a specific form for the energy generation rate, say

ǫnuc = ǫ0 ρTν (6.30)

so that eq. (6.11) can be written as

dldx= ǫ0 M ρTν (6.31)

By making use of the other homology relations, including themass-luminosity relation eq. (6.29),we obtain for a homogeneous, radiative star with constant opacity and consisting of an idealgas:

R∝ µν−4ν+3 M

ν−1ν+3 (6.32)

71

Page 20: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

The radius therefore depends onν, that is, on the mode of nuclear energy generation. Formain-sequence stars there are two possibilities:

pp-chain (ν ≈ 4) R∝ M0.43

CNO cycle (ν ≈ 18) R∝ µ2/3M0.81

The mass-luminosity and mass-radius relations (6.29) and (6.32) can be compared to the observedrelations for main-sequence stars that were presented in Chapter 1, and to the results of detailed stellarstructure calculations. We can also obtain homology relations for the central temperatures and centraldensities of main-sequence stars. This comparison is deferred to Chapter 8, where the main sequenceis discussed in more detail.

6.5.2 Homologous contraction

We have seen in Chapter 2 that, as a consequence of the virial theorem, a star without internal energysources must contract under the influence of its own self-gravity. Suppose that this contraction takesplace homologously. According to eq. (6.14) each mass shellinside the star then maintains the samerelative radiusr/R, so that

r(m)r(m)

=RR.

Since in this case we compare homologous models with the samemassM, we can replacex by themass coordinatem. For the change in density we obtain from eq. (6.17) that

ρ(m)ρ(m)

= −3RR, (6.33)

and if the contraction occursquasi-staticallythen the change in pressure follows from eq. (6.21),

P(m)P(m)

= −4RR=

43ρ(m)ρ(m)

. (6.34)

To obtain the change in temperature for a homologously contracting star, we have to consider theequation of state. Writing the equation of state in its general, differential form eq. (3.49) we caneliminateP/P to get

TT=

1χT

(43− χρ

)

ρ

ρ=

1χT

(3χρ − 4)RR. (6.35)

Hence the temperature increases as a result of contraction as long asχρ < 43. For an ideal gas,

with χρ = 1, the temperature indeed increases upon contraction, in accordance with our (qualitative)conclusion from the virial theorem. Quantitatively,

TT=

13ρ

ρ.

However, for a degenerate electron gas withχρ = 53 eq. (6.35) shows that the temperature decreases,

in other words a degenerate gas sphere willcool upon contraction. The full consequences of thisimportant result will be explored in Chapter 7.

72

Page 21: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

6.6 Stellar stability

We have so far considered stars in both hydrostatic and thermal equilibrium. But an important ques-tion that remains to be answered is whether these equilibriaarestable. From the fact that stars canpreserve their properties for very long periods of time, we can guess that this is indeed the case. Butin order to answer the question of stability, and find out under what circumstances stars may becomeunstable, we must test what happens when the equilibrium situation is perturbed: will the perturba-tion be quenched (stable situation) or will it grow (unstable situation). Since there are two kinds ofequilibria, we have to consider two kinds of stability:

• dynamical stability: what happens when hydrostatic equilibrium is perturbed?

• thermal (secular) stability: what happens when the thermalequilibrium situation is perturbed?

6.6.1 Dynamical stability

The question of dynamical stability relates to the responseof a certain part of a star to a perturbationof the balance of forces that act on it: in other words, a perturbation of hydrostatic equilibrium. Wealready treated the case of dynamical stability tolocal perturbations in Sec. 4.4.1, and saw that in thiscase instability gives rise toconvection. In this section we look at the global stability of concentriclayers within a star to radial perturbations, i.e. compression or expansion. A rigorous treatment ofthis problem is very complicated, so we will only look at a very simplified example to illustrate theprinciples.

Suppose a star in hydrostatic equilibrium is compressed on ashort timescale,τ ≪ τKH, so thatthe compression can be considered as adiabatic. Furthermore suppose that the compression occurshomologously, such that its radius decreases fromR to R′. Then the density at any layer in the starbecomes

ρ→ ρ′ = ρ(

R′

R

)−3

and the new pressure after compression becomesP′, given by the adiabatic relation

P′

P=

(

ρ′

ρ

)γad

=

(

R′

R

)−3γad

.

The pressure required for HE after homologous contraction is(

P′

P

)

HE=

(

ρ′

ρ

)4/3

=

(

R′

R

)−4

Therefore, ifγad >43 thenP′ > P′HE and the surplus pressure leads to re-expansion (on the dynamical

timescaleτdyn) so that HE is restored. If, however,γad <43 thenP′ < P′HE and the increase of pressure

is not sufficient to restore HE. The compression will therefore reinforce itself, and the situation isunstable on the dynamical timescale. We have thus obtained acriterion fordynamical stability:

γad >43 (6.36)

It can be shown rigorously that a star that hasγad >43 everywhere is dynamically stable, and ifγad =

43

it is neutrally stable. However, the situation whenγad <43 in some part of the star requires further

investigation. It turns out that global dynamical instability is obtained when the integral∫

(

γad−43

)Pρ

dm

73

Page 22: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

over the whole star is negative. Therefore ifγad <43 in a sufficiently large core, whereP/ρ is high,

the star becomes unstable. However ifγad <43 in the outer layers whereP/ρ is small, the star as a

whole need not become unstable.

Cases of dynamical instability

Stars dominated by an ideal gas or by non-relativistic degenerate electrons haveγad =53 and are

therefore dynamically stable. However, we have seen that for relativistic particlesγad→ 43 and stars

dominated by such particles tend towards a neutrally stablestate. This is the case ifradiation pressuredominates (at highT and lowρ), or if degenerate electrons become extremely relativistic (at very highρ). A small disturbance of such a star could either lead to a collapse or an explosion.

A process that can lead toγad <43 is partial ionization(e.g. H↔ H+ + e−), as we have seen in

Sec. 3.5. Since this normally occurs in the very outer layers, at smallP/ρ, it does not lead to overalldynamical instability of the star (although it is connectedto driving oscillations in some kinds ofstars).

However at very high temperatures two other processes can occur that have a similar effect toionization. These arepair creation(γ+γ ↔ e++e−, see Sect. 3.6.2) andiron disintegration(γ+Fe↔α). These processes, that may occur in massive stars in late stages of evolution, also lead toγad <

43

but now in the core of the star. These processes can lead to a stellar explosion or collapse (seeChapter 12).

6.6.2 Secular thermal stability

This question of thermal orsecularstability, i.e. the stability of thermal equilibrium, is intimatelylinked to the virial theorem. In the case of an ideal gas the virial theorem (Sect. 2.3) tells us that thetotal energy of a star is

Etot = −Eint =12Egr. (6.37)

which is negative: the star is bound. The rate of change of theenergy is given by the differencebetween the rate of nuclear energy generation in the deep interior and the rate of energy loss in theform of radiation from the surface:

Etot = Lnuc− L (6.38)

In a state of thermal equilibrium,L = Lnuc andEtot remains constant. Consider now a small pertur-bation of this situation, for instanceLnuc > L because of a small temperature fluctuation. This leadsto an increase of the total energy,δEtot > 0, and since the total energy is negative, its absolute valuebecomes smaller. The virial theorem, eq. (6.37), then tellsus that (1)δEgr > 0, in other words the starwill expand (δρ < 0), and (2)δEint < 0, meaning that the average temperature will decrease (δT < 0).Since the nuclear energy generation rateǫnuc ∝ ρTν depends on positive powers ofρ and especiallyT, the total nuclear energy generation will decrease:δLnuc < 0. By eq. (6.38) the perturbation toEtot

will be quenched and the state of thermal equilibrium will berestored. Thisthermostat, provided bythe virial theorem, is the mechanism that keeps stars in a stable state of thermal equilibrium for suchlong time scales.

We can generalise this to the case of stars with appreciable radiation pressure. For a mixture ofideal gas and radiation we can write, with the help of eqs. (3.13) and (3.14),

Pρ=

Pgas

ρ+

Prad

ρ= 2

3ugas+13urad. (6.39)

74

Page 23: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Applying the virial theorem in its general form, eq. (2.22),this yields

2Eint,gas+ Eint,rad = −Egr (6.40)

and the total energy becomes

Etot = −Eint,gas=12(Egr + Eint,rad). (6.41)

The radiation pressure thus has the effect of reducing the effective gravitational potential energy. Ifβ = Pgas/P is constant throughout the star, then eq. (6.41) becomes

Etot =12βEgr (6.42)

This is negative as long asβ > 0. The analysis of thermal stability is analogous to the caseof an idealgas treated above, and we see that stars in which radiation pressure is important, but not dominant,are still secularly stable. However, ifβ→ 0 then the thermostatic effect no longer works.

Thermal instability of degenerate gases

The arguments given above for secular stability depend crucially on the thenegative heat capacityofstars: the property that an increase of the total energy content leads to adecreaseof the temperature.This in turn depends on the proportionality of the internal energy to the temperature for an ideal gas.

In the case of a degenerate electron gas, the virial theorem eq. (6.37) still holds as long as theelectrons are non-relativistic – you may recall from Sect. 2.3 that the only assumption that went intoderiving this equation was thatu = 3

2P/ρ. However, for a degenerate gas the pressure and thereforethe internal energy is independent of the temperature (Sect. 3.3.5). If the same perturbationLnuc > Ldiscussed above is applied to a degenerate gas, the resulting decrease in internal energy will still leadto expansion (δρ < 0), but it will not lead to a decrease in temperature – instead, the temperaturewillincrease. This can be seen from eq. (6.35), valid in the case of homologous expansion, which we canwrite as

δTT=

1χT

(43− χρ

)

δρ

ρ. (6.43)

For a degenerate electron gas,χρ = 53 and 0< χT ≪ 1, henceδT > 0 upon expansion. Because

of the strong sensitivity of the nuclear energy generation rate toT, the perturbation will lead to anincrease ofLnuc, and thermal equilibrium will not be restored. Instead, thetemperature will continueto rise as a result of the increased nuclear energy release, which in turn leads to further enhancementof the energy generation. This instability is called athermonuclear runaway, and it occurs whenevernuclear reactions ignite in a degenerate gas. In some cases it can lead to the explosion of the star,although a catastrophic outcome can often be avoided when the gas eventually becomes sufficientlyhot to behave as an ideal gas, for which the stabilizing thermostat operates. From eq. (6.43) we seethat this happens when the value ofχρ drops below the critical value of43. In that caseδT changessign and becomes negative upon further expansion.

We shall encounter several examples of thermonuclear runaways in future chapters. The mostcommon occurrence is the ignition of helium fusion in stars with masses below about 2M⊙ – thisphenomenon is called thehelium flash. Thermonuclear runaways also occur when hydrogen gasaccumulates on the surface of a white dwarf, giving rise to so-callednova outbursts.

75

Page 24: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

The thin shell instability

In evolved stars, nuclear burning can take place in a shell around an inert core. If such a burningshell is sufficiently thin the burning may become thermally unstable, even under ideal-gas conditions.We can make this plausible by considering a shell with mass∆m inside a star with radiusR, locatedbetween a fixed inner boundary atr0 and outer boundary atr, so that its thickness isd = r − r0 ≪ R.If the shell is in thermal equilibrium, the rate of nuclear energy generation equals the net rate of heatflowing out of the shell (eq. 6.11). A perturbation by which the energy generation rate exceeds therate of heat flow leads to expansion of the shell, pushing the layers above it outward (δr > 0). Thisleads to a decreased pressure, which in hydrostatic equilibrium is given by eq. (6.34),

δPP= −4

δrr. (6.44)

The mass of the shell is∆m = 4πr02ρd, and therefore the density varies with the thickness of the

shell as

δρ

ρ= −δd

d= −δr

rrd. (6.45)

Eliminating δr/r from the above equations yields a relation between the changes in pressure anddensity,

δPP= 4

drδρ

ρ. (6.46)

Combining with the equation of state in its general, differential form eq. (3.49) we can eliminateδP/Pto obtain the resulting change in temperature,

δTT=

1χT

(

4dr− χρ

)

δρ

ρ. (6.47)

The shell is thermally stable as long as expansion results ina drop in temperature, i.e. when

4dr> χρ (6.48)

sinceχT > 0. Thus, for a sufficiently thin shell a thermal instability will develop. (In the case of anideal gas, the condition 6.48 givesd/r > 0.25, but this is only very approximate.) If the shell is verythin, the expansion does not lead to a sufficient decrease in pressure to yield a temperature drop, evenin the case of an ideal gas. This may lead to a runaway situation, analogous to the case of a degenerategas. The thermal instability of thin burning shells is important during late stages of evolution of starsup to about 8M⊙, during theasymptotic giant branch.

76

Page 25: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Chapter 7

Schematic stellar evolution –consequences of the virial theorem

7.1 Evolution of the stellar centre

We will consider the schematic evolution of a star, as seen from its centre. The centre is the pointwith the highest pressure and density, and (usually) the highest temperature, where nuclear burningproceeds fastest. Therefore, the centre is the most evolvedpart of the star, and it sets the pace ofevolution, with the outer layers lagging behind.

The stellar centre is characterized by the central densityρc, pressurePc and temperatureTc andthe composition (usually expressed in terms ofµ and/or µe). These quantities are related by theequation of state (EOS). We can thus characterize the evolution of a star by an evolutionary track inthe (Pc, ρc) diagram or the (Tc, ρc) diagram.

7.1.1 Hydrostatic equilibrium and the Pc-ρc relation

Consider a star in hydrostatic equilibrium (HE), for which we can estimate how the central pres-sure scales with mass and radius from the homology relations(Sec. 6.5). For a star that expands orcontracts homologously, we can apply eq. (6.23) to the central pressure and density to yield

Pc = C · GM2/3ρc4/3 (7.1)

whereC is a constant. This is a fundamental relation for stars in HE:in a star of mass M that expandsor contracts homologously, the central pressure varies as central density to the power43. The valueof the constantC depends on the density distribution in the star. However, itcan be shown thatthis dependence is only very weak. E.g., for polytropic models with indexn = 1.5 . . . 3 (this rangeencompasses most actual stars),C varies between 0.48. . . 0.36. Hence relation (7.1) is reasonablyaccurate, even if the contraction is not exactly homologous. In other words: for a star of a certainmass, the central pressure is almost uniquely determined bythe central density.1

1Equation (7.1) can also be derived from the virial theorem (VT) which, as we have seen, holds for all stars in HE andis in a sense equivalent to HE. In its most general form the VT can be written as

∫ M

0

dm= −13

Egr

which can easily be shown to lead to the same scaling of the average pressureP with M andρ as in equation (6.23). Sincefor homologous changesPc ∝ P andρc ∝ ρ we again obtain eq. (7.1).

77

Page 26: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Note that we have obtained this relation without considering the EOS. Therefore eq. (7.1) defines auniversal relation for stars in HEthat is independent of the equation of state. It expresses the fact thata star that contracts quasi-statically must achieve a higher internal pressure to remain in hydrostaticequilibrium. Eq. (7.1) therefore defines anevolution trackof a slowly contracting (or expanding) starin thePc-ρc plane.

7.1.2 The equation of state and evolution in thePc-ρc plane

By considering the EOS we can also derive the evolution of thecentral temperature. This is obviouslycrucial for the evolutionary fate of a star because e.g. nuclear burning requiresTc to reach certain(high) values. We start by considering lines of constantT, isotherms, in the (P, ρ) plane.

We have encountered various regimes for the EOS:

• Radiation dominated:P = 13aT4. Hence an isotherm in this region is also a line of constantP.

• (Classical) ideal gas:P =RµρT. Hence an isotherm hasP ∝ ρ.

• Non-relativistic electron degeneracy:P = KNR(ρ/µe)5/3 (eq. 3.37). This is independent ofT. More accurately: the complete degeneracy implied by this relation is only achieved whenT → 0, so this is in fact the isotherm forT = 0 (andρ not too high).

• Extremely relativistic electron degeneracy:P = KER(ρ/µe)4/3 (eq. 3.39). This is the isothermfor T = 0 and very highρ.

Figure 7.1 shows various isotherms schematically in the logρ - logP plane. Where radiationpressure dominates (lowρ) the isotherms are horizontal and where ideal-gas pressuredominates theyhave a slope= 1. The isotherm forT = 0, corresponding to complete electron degeneracy, has slope53 at relatively low density and a shallower slope4

3 at high density. The region to the right is forbiddenby the Pauli exclusion principle (electrons are fermions).

The dashed lines are schematic evolution tracks for stars ofdifferent massesM1 andM2. Accord-ing to eq. (7.1) they have a slope4

3 and the track for higher mass lies above that for lower mass.

Figure 7.1. Schematic evolution in the logρ - logP plane. See the text for an explanation.

78

Page 27: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Several important conclusions can be drawn from this diagram:

• As long as the gas is ideal, contraction (increasingρc) leads to a higherTc, because the slopeof the evolution track is steeper than that of ideal-gas isotherms: the evolution track crossesisotherms of higher and higher temperature.

Note that this is consistent with what we have already concluded from the virial theorem for anideal gas:Ein = −1

2Egr. Contraction (decreasingEgr, i.e. increasing−Egr) leads to increasinginternal (thermal) energy of the gas, i.e. to heating of the stellar gas!

• Tracks for masses lower than some critical valueMcrit (e.g. the track labeledM1) eventuallyrun into the line for complete electron degeneracy, becausethis has a steeper slope. Hence forstars withM < Mcrit there exist a maximum achievable central density and pressure, ρmax andPmax, which define the endpoint of their evolution. This endpointis a completely degeneratestate, i.e. a white dwarf, where the pressure needed to balance gravity comes from electronsfilling the lowest possible quantum states.

Because complete degeneracy corresponds toT = 0, it follows that the evolution track mustintersect each isotherm twice. In other words, stars withM < Mcrit also reach a maximum tem-peratureTmax, at the point where degeneracy starts to dominate, after which further contractionleads to decreasingTc. ρmax, Pmax andTmax all depend onM and increase with mass.

• Tracks for masses larger thanMcrit (e.g. the one labeledM2) miss the completely degenerateregion of the EOS, because at highρ this has the same slope as the evolution track. Fromthe physical point of view, the electrons become relativistic and their velocity cannot exceedc, so that they exert less pressure than if there were no limit to their velocity. Hence, electrondegeneracy is not sufficient to counteract gravity, and a star withM > Mcrit must keep oncontracting and getting hotter indefinitely (up to the pointwhere the assumptions we havemade break down, e.g. whenρ becomes so high that the protons inside the nuclei capture freeelectrons and a neutron gas is formed, which can also become degenerate).

Hence the evolution of stars withM > Mcrit is qualitatively different from that of stars withM < Mcrit. This critical mass is none other than theChandrasekhar massthat we have alreadyencountered (K&W section 19.7). It is the unique mass of a completely degenerate and extremelyrelativistic gas sphere (MCh = 1.46M⊙ for a composition of He or heavier elements withµe = 2). Astar withM = MCh must also collapse under its own gravity but the electrons become degenerate andextremely relativistic in the process.

7.1.3 Evolution in theTc-ρc plane

We now consider how the stellar centre evolves in theTc, ρc diagram. First we divide theT, ρ planeinto regions where different processes dominate the EOS, see Sec. 3.3.7 and Fig. 3.4, reproducedbelow in Fig. 7.2a.

For a slowly contracting star in hydrostatic equilibrium equation (7.1) implies that, as long as thegas behaves like a classical ideal gas:

Tc ρc = C GM2/3ρc4/3 → Tc ∝ µM2/3ρc

1/3 (7.2)

(Compare to Sec. 6.5.2.) This defines an evolution track in the logT, logρ plane with slope13. Stars

with different mass evolve along tracks that lie parallel to each other, those with largerM lying athigher Tc and lowerρc. Tracks for larger mass therefore lie closer to the region where radiation

79

Page 28: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

−5 0 5 10 4

6

8

10

log ρc

log

T c

radiation

ideal gas degenerate

NR ER

−5 0 5 10 4

6

8

10

log ρc

log

T c

0.1

1.0

10

100

MCh

Figure 7.2. The equation of state in the logTc, logρc plane, with approximate boundaries between regionswhere radiation pressure, ideal gas pressure, non-relativistic electron degeneracy and extremely relativisticelectron degeneracy dominate (left panel). In the right panel, schematic evolution tracks for contracting starsof 0.1...100M⊙ have been added.

pressure is important:the larger the mass of a star, the more important is the radiation pressure.Furthermore, the relative importance of radiation pressure does not change as a star contracts, becausethe track runs parallel to the boundary between ideal gas andradiation pressure.2

As the density increases, stars withM < MCh approach the region where non-relativistic electrondegeneracy dominates, because the boundary between ideal gas and NR degeneracy has a steeperslope than the evolution track. Inside this region, equating relation (7.1) to the NR degenerate pressuregives:

K1ρc

1/3

µe5/3= C GM2/3 → ρc ∝ µe

5M2 (7.3)

When degeneracy dominates the track becomes independent ofTc, and the star moves down alonga track of constantρc. This is theρmax we found from thePc, ρc diagram. The larger the mass, thehigher this density. (When the electrons become relativistic atρc ∼> 106 g/cm3, the pressure increasesless steeply with density and the central density for a degenerate star of massM is in fact larger thangiven by eq. 7.3).

Equations (7.2) and (7.3) imply that, for a star withM < MCh that contracts quasi-statically,Tc

increases asρc1/3 until the electrons become degenerate. Then a maximum temperature is reached,

and subsequently the star cools at a constant density when degenerate electrons provide the pressure.The schematic evolution tracks for 0.1 and 1.0M⊙ given in Fig. 7.2 show this behaviour. Combiningeqs. (7.2) and (7.3) indicates that the maximum central temperature reached increases with stellarmass asTc ∝ M4/3 (see Exercise 8.2).

For M > MCh, the tracks of contracting stars miss the degenerate regionof theT-ρ plane, becauseat high density the boundary between ideal gas and degeneracy has the same slope as an evolution

2It is easy to show for yourself that the evolution track for a star in which radiation pressure dominates would have thesame slope of13 in the logT, logρ plane. However, such stars are very close to dynamical instability.

80

Page 29: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

track. The pressure remains dominated by an ideal gas, andTc keeps increasing likeρc1/3. This

behaviour is shown by the schematic tracks for 10 and 100M⊙.

7.2 Nuclear burning regions and limits to stellar masses

We found that stars withM < MCh reach a maximum temperature, the value of which increases withmass. This means that only gas spheres above a certain mass limit will reach temperatures sufficientlyhigh for nuclear burning. The nuclear energy generation rate is a sensitive function of the temperature:

ǫnuc = ǫ0 ρλTν (7.4)

where for most nuclear reactions (those involving two nuclei) λ = 1, while ν depends mainly on themasses and charges of the nuclei involved and usuallyν ≫ 1. For H-burning by the pp-chain,ν ∼ 4and for the CNO-cycle which dominates at somewhat higher temperature,ν ∼ 18. For He-burning bythe 3α reaction,ν ∼ 40 (andλ = 2 because three particles are involved). For C-burning and O-burningreactionsν is even larger. As a consequence of thisT-sensitivity:

• each nuclear reaction takes place at a particular nearly constant temperature, and

• nuclear burning cycles of subsequent heavier elements are well separated in temperature

As a star contracts and heats up, nuclear burning becomes important when the energy generated,Lnuc =

ǫnucdm, becomes comparable to the energy radiated away from the surface,L. From thismoment on, the star can compensate its surface energy loss bynuclear energy generation. The firstnuclear fuel to be ignited is hydrogen, which burns atT ∼ 107 K.

We can estimate the minimum mass required for hydrogen burning by comparing the maximumcentral temperature reached by a gas sphere of massM, to the central temperature at which hydrogenfusion takes place in a star of massM. An estimate of the latter can be obtained from main-sequencehomology relations (e.g. see Sec. 6.5.1). By doing this we find a minimum mass for H-burning of∼ 0.1 M⊙.

Detailed calculations reveal that H-burning is ignited in protostars withM > 0.08M⊙. Lessmassive objects become partially degenerate before the required temperature is reached and continueto contract and cool without ever burning hydrogen. Such objects are not stars according to ourdefinition (Ch. 1) but are known asbrown dwarfs.

We have seen earlier that the contribution of radiation pressure increases with mass, and becomesdominant forM ∼> 100M⊙. A gas dominated by radiation pressure has an adiabatic index γad =

43,

which means that hydrostatic equilibrium in such stars becomes dynamically unstable. The criterionfor dynamical stability isγad >

43 (see Sec. 6.6.1), which is satisfied for an ideal gas withγad =

53.

Therefore stars much more massive than 100M⊙ cannot be stable, and indeed none are known toexist (while those withM > 50M⊙ indeed show signs of marginal stability, e.g. they lose massveryreadily).

Hence stars are limited to a rather narrow mass range of∼ 0.1 M⊙ to ∼ 100M⊙. The lowerlimit is set by the minimum temperature required for nuclearburning, and the upper limit by therequirement of dynamical stability.

81

Page 30: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

7.2.1 Overall picture of stellar evolution and nuclear burning cycles

As a consequence of the virial theorem, a self-gravitating gas sphere in HE must contract and heat up(assuming the gas is ideal) when it radiates energy from the surface:

L = −Etot = Ein = −12Egr ≈

Egr

τKH(7.5)

This is the case for protostars that have formed out of an interstellar gas cloud. The evolution, i.e.overall contraction, takes place on a thermal timescaleτKH. As the protostar contracts and heatsup, the nuclear energy generation rate (which is at first negligible) increases in the centre, until theburning rate matches the energy loss from the surface:

L = −Enuc ≈Enuc

τnuc(7.6)

At this point, contraction stops andTc andρc remain approximately constant, at the values needed forhydrogen burning. The stellar centre occupies the same place in theTc-ρc diagram for about a nucleartimescaleτnuc.

When H is exhausted in the core (which has a mass typically∼ 0.1 of the total massM, and nowconsists of helium) this He core resumes its contraction. Meanwhile the layers around it expand. Thisconstitutes a large deviation from homology and relation (7.1) no longer applies to the whole star.However the core itself still contracts more or less homologously, while the weight of the envelopedecreases as a result of its expansion. Therefore relation (7.1) is approximately valid for thecoreofthe star, if we replaceM by the core massMc. The core continues to contract and heat up at a paceset by its own thermal timescale,

Lcore≈ Ein,core≈ −12Egr,core≈

Egr,core

τKH,core(7.7)

as long as it behaves as an ideal gas. It is now the He core mass (rather than the total mass of the star)that determines the further evolution.

−5 0 5 10 4

6

8

10

log ρc

log

T c

radiation

ideal gas degenerate

NR ER

pp

CNOH-b

He-b

C-bO-b

Si-b

−5 0 5 10 4

6

8

10

log ρc

log

T c

H-b

He-b

C-bO-b

Si-b

0.1

1.0

10

100

Figure 7.3. The equation of state and regions of nuclear burning (left) and schematic evolution tracks forcontracting stars of 0.1...100M⊙ (right) in the logTc, logρc plane.

82

Page 31: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Arguments similar to those used for deriving the minimum mass for H-burning, lead to a minimumrequired (core) mass for He-ignition, which according to detailed calculations is≈ 0.3 M⊙. Stars witha core mass larger than this value ignite He in the centre whenTc ≈ 108 K, which stops furthercontraction while the energy radiated away can be supplied by He-burning reactions. This can go onfor a nuclear timescale of He burning, which is about 0.1 times that of H burning. In stars with a Hecore mass< 0.3 M⊙ the core becomes degenerate before reachingTc = 108 K, and in the absenceof a surrounding envelope it would cool to become a white dwarf composed of helium. (In practice,however, H-burning in a shell around the core keeps the core hot and whenMc has grown to≈ 0.5 M⊙He ignites in a degenerate flash.)

Following a similar line of reasoning, the minimum (core) mass for C-burning (requiringT ≈5× 108 K) is ≈ 1.1 M⊙. Less massive cores are destined to never ignite carbon and become CO whitedwarfs. The minimum core mass required for the next stage, Ne-burning, turns out to be≈ MCh.Stars that develop cores withMc > MCh therefore also undergo all subsequent nuclear burning stages(Ne-, O- and Si-burning) because they never become degenerate and continue to contract and heatafter each burning phase. Eventually they develop a core consisting of Fe, from which no furthernuclear energy can be squeezed. The Fe core must collapse in acataclysmic event (a supernova or agamma-ray burst) and become a neutron star or black hole.

Fig. 7.3 shows a schematic picture of the evolution of stars of different masses in theT, ρ diagram,while Fig. 7.4 presents the result of detailed calculationsfor various masses.

To summarize, we have obtained the following picture. Nuclear burning cycles can be seen as long-lived but temporary interruptions of the inexorable contraction of a star (or at least its core) underthe influence of gravity. This contraction is dictated by thevirial theorem, and a result of the factthat stars are hot and lose energy by radiation. If the core mass is less than the Chandrasekhar mass,then the contraction can eventually be stopped (after one ormore nuclear cycles) when electrondegeneracy supplies the pressure needed to withstand gravity. However if the core mass exceedsthe Chandrasekhar mass, then degeneracy pressure is not enough and contraction, interrupted bynuclear burning cycles, must continue at least until nuclear densities are reached.

83

Page 32: Chapter 5 - uni-bonn.denlanger/siu_web/ssescript/chapter5-7.pdf · Chapter 5 Nuclear reactions in stars The contents of this chapter are covered by Chapter 18 of K & W . What follows

Figure 7.4. Detailed evolution tracks in the logρc - logTc plane for masses between 1 and 15M⊙. Theinitial slope of each track (labelled pre-main sequence contraction) is equal to1

3 as expected from our simpleanalysis. When the H-ignition line is reached wiggles appear in the tracks, because the contraction is then nolonger strictly homologous. A stronger deviation from homologous contraction occurs at the end of H-burning,because only the core contracts while the outer layers expand. Accordingly, the tracks shift to higher densityappropriate for their smaller (core) mass. These deviations from homology occur at each nuclear burningstage. Consistent with our expectations, the most massive star (15M⊙) reaches C-ignition and keeps evolvingto higherT andρ. The core of the 7M⊙ star crosses the electron degeneracy border (indicated byǫF/kT = 10)before the C-ignition temperature is reached and becomes a C-O white dwarf. The lowest-mass tracks (1 and2 M⊙) cross the degeneracy border before He-ignition because their cores are less massive than 0.3M⊙. Basedon our simple analysis we would expect them to cool and becomeHe white dwarfs; however, their degenerateHe cores keep getting more massive and hotter due to H-shell burning. They finally do ignite helium in anunstable manner, the so-called He flash.

84