Stars and Planets

< < Module 5 > >

—Stellar Evolution—

Chris Ormel

Roadmap module 5

Opacity

The "resistance" light experiences in penetrating matter

(→ M4. Eddington Luminosity M6. Atmospheres)

Heat transport

Mechanism to transport heat (radiative diffusion, convection, conduction)

 

Stellar structure equations

The equations that govern the structure and evolution of stars. Usually solved numerically with codes like MESA

|

Homology

A simplification to the stellar structure equations that allows for analytical scaling solutions

Stellar evolution

A description of how stars interior and appearance changes with time and how these changes depend on parameters, such as the initial mass

High-mass stellar evolution

High-mass stars deviate from evolution from low-mass stars and play a critical role in nucleosynthesis

  • radiative and convective temperature gradient
  • Schwarzschield stability criterion
  • mass-radius and mass-luminosity relationships
  • Hayashi track
  • H (shell) burning
  • He flash
  • Stellar nucleosynthesis

Topics

  • Random walk & opacity
  • Radiative and convective heat transport
  • Stellar structure equations
  • Homology

Mean free path and opacity

— read CO Ch. 9.2 and 9.3 —

 

Examples of three random walks. The mean free path is 10% of the radius of the circle. Eventually the photon escapes (statistically after 102 random walks).
  • (photon) mean free path : distance before a photon gets absorbed or scattered:

    ni: number density of absorbers per unit volume;
    σ: absorption cross section

  • opacity : inverse mean free path — the transparancy — normalized by density

    Other definitions for opacity exist. In our definition it is the total cross section per unit mass of the medium (units: cm2 g-1).

  • optical depth : number of mean free path length over a length L

Mean free path and opacity

— read CO Ch. 9.2 and 9.3 —

 

the optical depth is depth normalized by the mean free path of the medium.
  • (photon) mean free path : distance before a photon gets absorbed or scattered:

    ni: number density of absorbers per unit volume;
    σ: absorption cross section

  • opacity : inverse mean free path — the transparancy — normalized by density

    Other definitions for opacity exist. In our definition it is the total cross section per unit mass of the medium (units: cm2 g-1).

  • optical depth : number of mean free path length over a length L

Mean free path and opacity

— read CO Ch. 9.2 and 9.3 —

 

only radiation below τ~1 will emerge
  • (photon) mean free path : distance before a photon gets absorbed or scattered:

    ni: number density of absorbers per unit volume;
    σ: absorption cross section

  • opacity : inverse mean free path — the transparancy — normalized by density

    Other definitions for opacity exist. In our definition it is the total cross section per unit mass of the medium (units: cm2 g-1).

  • optical depth : number of mean free path length over a length L

opacity

— see CO 9.2 —

 

Sources of opacity:

  • electron scattering

    where is the Thomson cross section (see module 4).

    • Electron scattering dominates in stellar interiors. where gas is ionized.
    • At very high T (frequency), we have that (transition to Compton scattering)

Question: Why can't free electrons absorb photons?

Question: Why the (1+X) factor?

photon scattering and absorption: bound-free, bound-bound, free-free, and electron scattering. In stellar interiors, electron scattering determines the opacity.

opacity

— see CO 9.2 —

 

Sources of opacity:

  • electron scattering
  • Free-free and bound-free

    (cgs units). These forms are Kramer's laws where frequency-dependent opacities have been averaged over the radiation field (Rosseland mean). Expressions are very approximate and are valid only in a certain T-range.

Question: How do we get the combined opacity at a single frequency, κν?

Question: Why is the Rosseland mean opacity harmonically averaged over the radiation field?

bound-free opacities for the H-atom

These opacities are so called Rosseland mean opacities:

that is, the inverse of the opacity (the "transparancy") is weighted by the radiation field.

Why the suddon drop in opacity to 0 at these "cliffs"?

opacity

— see CO 9.2 —

 

Sources of opacity:

  • electron scattering
  • Free-free and bound-free
  • negative H (H) opacity

    valid for 3,000 < T < 6,000 and -10 < log10 ρ < -5. Important in "cool" atmospheres, where neutral H can bind a second electron. This electron has a low binding energy (compared to neutral H); it dissociates with photons energies above 0.75 eV (~104 K).

  • conductive opacities

    degenerate regions

bound-free opacities for the H-atom
HI bound-free extinction coefficient per hydrogen atom in level n. These must be weighted by the level population and added. (c) Gray (1992).

These opacities are so called Rosseland mean opacities:

that is, the inverse of the opacity (the "transparancy") is weighted by the radiation field.

solar opacities

 

use opacity tables to find as function of temperature and pressure.

We identify with increasing T and ρ:

  1. H opacity near the Sun's photosphere
  2. strong increase in H opacity. After this peak H-ionization reduces the H opacity
  3. bound-free opacities, followed by free-free dominate in the interior
  4. solar core. Opacities just higher than electron scattering limit
a
b
c
d
Radiative opacity for stars, https://opalopacity.llnl.gov/existing.html

Radiative and convective heat transport

— read CO 10.4 —

 

Energy transport by diffusion of radiation results in an energy flux (units: [E L-2 T-1]) of

with the radiation energy density, the photon diffusivity, the mean free path, c the speed of light, the opacity. Equating this to the luminosity flux, one obtains

where is the opacity weighted by the radiation field (Rosseland mean). Following convention, the temperature gradient is recast in terms of a gradient with respect to pressure

With the hydrostatic balance equation, this radiative temperature gradient becomes

needs to adjust to in order for the energy to be transported by radiation!

spread of a dye dissolved in water. Diffusion acts to make the concentration of the uniform, which results in a flux of the dye towards regions of low concentration (down). (c) wikipedia

Radiative and convective heat transport

 

The stability condition against convection is

the adiabatic exponent. A strong density gradient (stratification) — dense material below light material — stabilizes fluids.

If we define the EoS as:

where for an ideal EoS. In the adiabatic case, we obtain

the stability criterion transforms to:

mixing length
lighter blobs
rise
deposite energy
after lmix
Convective heat transport: blobs are lighter than the environment, causing them to rise to deposit their energies in cooler layers. Image source: wikipedia

Radiative and convective heat transport

 

some thermodynamical relations may be helpful

where are the heat capacities at constant pressure and temperature. from this the heat capacity ratio follows. Another is the first law of thermodynamics:

if we put we obtain a relationship between the heat capacity ratios

mixing length
lighter blobs
rise
deposite energy
after lmix
Convective heat transport: blobs are lighter than the environment, causing them to rise to deposit their energies in cooler layers. Image source: wikipedia

Radiative and convective heat transport

 

When the radiative gradient becomes too steep, convective "blobs" will rise because they stay lighter than the surroundings. The condition for convection is:

which is Schwarzschield criterion (ignoring any composition gradient). More preciese, in a convective medium, the relation

holds.

  • the actual temperature gradient is less than the radiative gradient as part of the flux is carried by convection.
  • must hold for convection
  • Blobs themselves are never fully adiabatic ( ).

Theories as Mixing length theory offer a phenomenological model to calculate the energy flux that can be carried by convection. It turns out that that under most (stellar) conditions is an acceptable approximation. In other words, convection in stars is very effective. The temperature gradient only needs to be slightly above adiabatic to carry the entire energy flux!

Schematic for the temperature gradient

Radiative and convective heat transport

 

In mixing length theory the convective blobs are assumed to travel for a mixing length before equilibriating with the surroundings

where δq and δT are differences betweeen the heat contents and temperature of the blobs with its surroundings, e.g., cP is the heat capacity at constant pressure, vbl is the velocity of the blobs. After further manipulation and phenomenological reasoning, one obtains

where α ~ β ~ 1 and If we take (the convective bubbles are anyway modelled as adiabatic), it follows that for stellar conditions only a slight overadiabicity, suffices to transport the energy flux by convection.

mixing length
lighter blobs
rise
deposite energy
after lmix
Convective heat transport: blobs are lighter than the environment, causing them to rise to deposit their energies in cooler layers. Image source: wikipedia

Stellar structure equations:

read CO p.330—334

Stellar structure equations

  • Lagrangian formulation is used; radius coordinates r exchanged for mass m
  • the temperature gradient must be provided. Very often

    which is the Schwarzschield criterion. But other choices (e.g. Ledoux criterion) are available.

where m is mass, r is radius, P pressure, T temperature, l luminosity, ρ density, G Newton's gravitational constant the temperature gradient, s entropy, ε the energy production rate, and Xi the composition.

Stellar structure equations:

The stellar structure equations represent a system of ordinary differential equations (ODEs). To solve them, we need boundary conditions:

  • at the center, m=0: L=r=0 (T and P unknown)
  • at the surface, m=M, we can take zero boundary conditions for temperature and pressure: T=P=0
  • better to take photospheric boundary conditions:

    These relations follow from the Eddington approximation for the atmosphere:

    at τ=2/3 the photosphere the temperature equals the effective temperature and the pressure can be found by integrating dP/dr to infinity

Schematic for solving the stellar structure equations. The four first order ODE must be supplemented by four boundary conditions. Unknowns are in red. Typically atmospheric boundary conditions are used.

Stellar evolution — A numerical intermezzo

Stellar evolution is a thoroughly numerical problem. Everything we "know" about star's interiors comes from numerical modelling. Yet, it has proven very successful. Overall, there is a very precise understanding about stellar evolution — especially compared to other astrophysical research fields.

MESA is a publicly available collection of codes ("modules"), aimed towards 1D stellar evolution, which are contributed by many users from the entire planet!

  • it is one dimension (1D), so relatively fast
  • illuminating and rich
  • large international community


Above all, it's fun! Strongly recommend the ambitious student to download and check it out!

MESA — Modules for Experiments in Stellar Astrophysics
Q: What is the name of this diagram?
Q: t=1—10 Gyr: Why does MESA evolve
this phase so efficiently?
Q: t=1.2 Gyr: L increases.
What is the name of this phase?
Q: RBG phase:
Why is there no H-burning in the core?
Q: What happens at the end?
example of a 1 solar mass model run of MESA by your instructor.

Elements of Stellar Evolution

read CO Ch.13.1—13.2

 

For a solar mass star: 1 molecular cloud collapse; 2 fragmentation; 3 contraction and heat up; 4 protostars and disks; 5 contraction slowing; 6 onset of fusion; 7 main sequence; 8 H shell burning; 9 He flash; 10 He burning; 11 double shell burning; 12 planetary nebular; 13 White dwarf; 14 WD cooling

Homology & stellar evolution

 

Some insight into the stellar structure equations can be obtained by comparing two stars of different mass, assuming homology

where is the homologous mass coordinate


Note: The conditions for homology are only rarely satisfied.

For example, when a stars' core become degenerate while the outer regions expands — the RGB stage — it breaks

Under the homology assumption stars of different mass (or one star at different times) share the same relative radius-mass relationship.

Homology & stellar evolution

 

Applying homology to the stellar structure equations, we obtain:

Homology relations obtained from the continuity equation, pressure balance, the (ideal) equation of state, and energy transport (by radiation) equation. LHS should be evaluated at the homologous position x . The latter relation shows that (for constant opacity and composition) , which is approximately the Main Sequence.

Under the homology assumption stars of different mass (or one star at different times) share the same relative radius-mass relationship.

Homology & stellar evolution

 

Applying the homologous relationship for the center ( ) to the general version of the EoS:

where we ignore a compositional gradient. Recall that
  • for the ideal EoS and
  • for a non-relativistic degnerate electron gas

Therefore, for a collapsing star:

Contracting (cooling!) stars heat up for an ideal gas, but contract for a degenerate stars!

Arrows denote direction into which a collapsing star evolves

Homology & stellar evolution

 

Stellar evolution can be understood as a continuous (quasi-static) collapse, interceded by prolonged periods of "steady" nuclear burning

  • low-mass stars do not reach H-burning temperatures

    They end up as Brown Dwarfs.

hydrogen burning
BD
BD
Temperature and density conditions apply to the core. All curves are indicative.

Homology & stellar evolution

 

Stellar evolution can be understood as a continuous (quasi-static) collapse, interceded by prolonged periods of "steady" nuclear burning

  • low-mass stars do not reach H-burning temperatures
  • solar-mass stars reach H-burning and He-burning temperatures, but cannot avoid degenerate cooling once their supply runs out

    Note: after H-burning the temperature in the now He-core would increase due to the higher μ. This is not reflected in the figure.

    Stars end up as degenerate objects (White Dwarfs)

hydrogen burning
He burning
WD
sun
Temperature and density conditions apply to the core. All curves are indicative.

Homology & stellar evolution

 

Stellar evolution can be understood as a continuous (quasi-static) collapse, interceded by prolonged periods of "steady" nuclear burning

  • low-mass stars do not reach H-burning temperatures
  • solar-mass stars reach H-burning and He-burning temperatures, but cannot avoid degenerate cooling once their supply runs out
  • high-mass stars avoid the "degeneracy trap", burn many elements

    More massive than the Chandrasekhar mass; They eventually explode as supernovae

hydrogen burning
He burning
C burning
O burning
Temperature and density conditions apply to the core. All curves are indicative.

Homology & stellar evolution

 

Stellar evolution can be understood as a continuous (quasi-static) collapse, interceded by prolonged periods of "steady" nuclear burning

  • low-mass stars do not reach H-burning temperatures
  • solar-mass stars reach H-burning and He-burning temperatures, but cannot avoid degenerate cooling once their supply runs out
  • high-mass stars avoid the "degeneracy trap", burn many elements
  • Even higher mass stars would be unstable due to radiation pressure
hydrogen burning
He burning
C burning
O burning
Temperature and density conditions apply to the core. All curves are indicative.

Evolution in the HR diagram — The Hayashi line

— read CO 12.3 —

  • In the limit of fully convective stars we can derive the following scaling relationships
    This line is known as the Hayashi line. These relationships follow from the following assumptions:
    • an n=3/2 polytropic (interior) model, , and .
    • an atmospheric model that matches the polytropic model (pressure equilibrium)
    • an opacity relation (H opacity)  , so and .
    • The Stefan-Boltzmann relationship (to eliminate R in liu of L)
  • The extremely weak dependence on T renders the line vertical in the HR-diagram

    (although more realistic models will feature some structure)

  • Stars in hydrostatic balance cannot exist in hydrostatic equilibrium right of this line (at lower T) !

    This is known as the "forbidden zone"


→ Pre main-sequence evolution tracks and observations in Taurus-Auriga star formation complex (Stahler 1988).

Hayashi
tracks
ZAMS
forbidden
region

 

The main sequence (MS)

— read CO 13.2 —

  • The place in the HR-diagram where stars burn ("fuse") hydrogen
  • Stars follow a mass-luminosity relationship with for low-mass MS-stars

    This means high-mass stars quickly leave the MS, while low-mass stars are still on the MS
    The main-sequence turnoff is a way to determine the age of globular cluster stars.

  • For stellar clusters, the MS-turnoff is an indicator of its age

Perryman et al. (1997) — Hertzsprung-Russell diagram for stars in the solar neighborhood, which distances have been determined by the Hipparcos spacecraft.

MS

 

The main sequence (MS)

  • The place in the HR-diagram where stars burn ("fuse") hydrogen
  • Stars follow a mass-luminosity relationship with for low-mass MS-stars

    This means high-mass stars quickly leave the MS, while low-mass stars are still on the MS
    The main-sequence turnoff is a way to determine the age of globular cluster stars.

  • For stellar clusters, the MS-turnoff is an indicator of its age

Question: What does the branch (A) below the MS represent?


Gaia Collaboration et al. (2018) — 20 years later, there is Gaia.

A

 

The main sequence (MS)

  • The place in the HR-diagram where stars burn ("fuse") hydrogen
  • Stars follow a mass-luminosity relationship with for low-mass MS-stars
  • For stellar clusters, the MS-turnoff is an indicator of its age

After the main sequence

  • Stars will ascend the Red Giant Branch and (later) the Asymptotic Giant Branch

    The evolution depends on stellar mass and also metallicity

  • The final outcome depends mainly on the (initial) stellar mass:
masscore-burningshell burningoutcome
~0.1H (convective)noneHe WD
~0.1—0.5HHH/He WD
~0.5—2.3H, HeHCO WD
~2.3—8H, HeH,HeC-rich WD
~6.0—8H, He,CH,HeO-rich WD
>8H, He,..., Sivariouscore collapse Supernovae

Question: These helium white dwarfs haven't been observed yet. Why not?

Evolutionary tracks for a solar-type star. (c) wikipedia

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

convection
nuclear
fusion

(B) hydrogen is exhausted in the center; onset of H-shell burning; growth of envelope.

The H-shell burning and the contraction of the (near-)degenerate He core, cause the envelope to expand!

The underlying reason is that nuclear reaction are very sensitive to temperature; a tiny increase in T (due to contraction of the shell), greatly increases the luminosity output

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

Q: Why does the convective zone grow during the RGB phase

convection
nuclear
fusion

(C) the expanding envelope becomes convective; the star moves up ("ascend") the Red Giant Branch (RGB).

The He-degenerate core contracts and grows; Luminosity and nuclear burning increase, resulting in higher mass (and smaller!) core; the high temperature gradient ensures that the nuclear burning shell is thin

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

convection
nuclear
fusion

(D-F) an accelerating evolution along the red giant branch.

The envelope has by now become largely convective (high luminosity and high opacity!); evolution along the Hayashi line

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

Q: Why doesn't the energy liberated in the He flash come out?

convection
nuclear
fusion

(F) He is (explosively) ignited in the degenerate core

A Helium flash occurs; the He-burning is (for a short period) out of control, reaching power levels comparable to the galaxy. However, most of the generated energy does not reach the surface. Because the degenerate core's pressure is insensitive to temperature. We do not observe this as the energy goes into internal energy, lifting the degeneracy!

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

convection
nuclear
fusion

(G-H) He core burning

He-core burning expands the core and causes the envelope to shrink. These He-burning stars occupy the so-called Horizontal branch

Evolution of a solar-mass star

 

Left: Example of a so-called Kippenhahn diagram — evolution of the internal physical structure with time. Red-shaded regions denote nuclear burning, grey convection. Right: corresponding curves in the HR-diagram. (c) Onno Pols

convection
nuclear
fusion

(H-J) He shell burning; the star ascends the Asymptotic Giant Branch (AGB)

On the AGB the star may experience strong pulsations, sheds its outer envelope to the ISM, before its He-fuel is exhausted. A degenerate C/O-core remains as a White Dwarf

High-mass star evolution

— read CO Ch. 15.3 —

  • In high-mass stars, a progression of burning cycles results in an onion-shell structure
  • Since Fe-burning would consume energy, the core collapses and heats up
  • Photons become energetic enough to disintegrate the Fe-group heavy nuclei (photo-disintegration)

    This reverses the nuclear fusion in a few seconds! For example:

  • Under the high density and temperatures, protons and electrons combine (electron capture)

    this releases a huge amount of neutrinos (cooling)

  • Collapse is halted by degenerate pressure from neutrons core-collapse (Type II) supernovae
(c) Cosmin Deaconu — onion shell structure of nuclear fusion shells. Timescales apply of a 25 solar mass star. Not to scale.

supernova in the Whirlpool galaxy

background image source: APOD

Supernovae 1987A

(c) Hubble Space Telescope/NASA, ESA; R. Kirshner, and M. Mutchler; R. Avila

Supernovae 1987A

(c) Hubble Space Telescope/NASA, ESA; R. Kirshner, and M. Mutchler; R. Avila

SN 1987A Is the most-studied supernovae. It exploded in 1987 in the Large Magellanic Cloud (LMC). See astronomy.com article

X-ray: NASA/CXC/U.Colorado/S.Zhekov et al.; Optical: NASA/STScI/CfA/P.Challis. See wikipedia article
(c) ALMA (ESO/NAOJ/NRAO), P. Cigan and R. Indebetouw; NRAO/AUI/NSF, B. Saxton; NASA/ESA

nucleosynthesis

— read CO p.542 —

Nucleosynthesis describes the process of producing the atomic nuclei. We distinghuish:

  • big-bang nucleosynthesis, responsible primarily for the Y=25% cosmic mass fraction of H-1, H-2, He-3, and He-4
  • stellar nucleosynthesis for the heavier elements by nuclear fusion

Elements heavier than Fe are produced by neutron (n) capture

  • s-process ("slow"): n-capture produces an unstable isotope and is followed by β-decay

    β-decay: (for a free neutron)

  • r-process ("rapid"): successive n-captures possible, due to extremely high neutron fluxes (as in core-collapse supernovae)
Isotope formation by s- and r-processes. Neutron capture increases the mass number A (in this figure the symbol N is used) but not the atomic number Z. Usually, unstable isotopes (light grey) decay by β-decay. But in core-collapse supernovae the neutron capture occurs faster than β-decay, resulting in the formation of elements by the rapid (r) process (indicated in blue). (c) Lamers & Levesque (2017)

end of module 5

—congrats—