Stellar Evolution
Stellar evolution is the process by which a star changes over the course of time.
Depending on the mass of the star, its lifetime can range from a few million years for the
most massive to trillions of years for the least massive, which is considerably longer
than the age of the universe.
Stars go through a lifecycle fueled by nuclear fusion. They begin in the main-sequence
phase, where hydrogen fuses in the core. As hydrogen depletes, stars expand into red
giants. Different masses lead to varied fusion processes, with larger stars forming
heavier elements. When a star exhausts its fuel, it collapses into a white dwarf,
expelling outer layers as a planetary nebula. Red dwarfs are predicted to brighten
before becoming low-mass white dwarfs. Stellar evolution is studied through
populations since changes occur too slowly to observe in individual stars.
The change in size with
time of a Sun-like star
Artist's depiction
of the life cycle of
a Sun-like star, starting as a main-sequence star at lower left then expanding through
the subgiant and giant phases, until its outer envelope is expelled to form a planetary
nebula at upper right.
Star Formation
Protostar
Stellar evolution starts with the collapse of a giant molecular cloud, leading to the
formation of protostars from smaller fragments. These protostars accrete gas and dust,
becoming pre-main-sequence stars. Their development depends on mass and is
influenced by gravitational forces and magnetic fields within dense filaments. Infrared
observations, like those from WISE, reveal protostars hidden in dust.
Brown dwarfs and sub-stellar objects
Protostars with masses less than roughly 0.08 M☉ (1.6×1029 kg) never reach
temperatures high enough for nuclear fusion of hydrogen to begin. These are known as
brown dwarfs. The International Astronomical Union defines brown dwarfs as stars
massive enough to fuse deuterium at some point in their lives (13 Jupiter masses (MJ),
2.5 × 1028 kg, or 0.0125 M☉). Objects smaller than 13 MJ are classified as sub-brown
dwarfs (but if they orbit around another stellar object they are classified as planets).
Both types, deuterium-burning and not, shine dimly and fade away slowly, cooling
gradually over hundreds of millions of years.
Main sequence stellar mass objects
More massive protostars initiate nuclear fusion at 10 million Kelvin, starting with the
proton-proton chain and progressing to helium. Stars slightly over 1 solar mass utilize
carbon-nitrogen-oxygen fusion (CNO cycle). Fusion establishes hydrostatic equilibrium,
balancing gravitational forces and initiating the main-sequence phase. A star's position
on the Hertzsprung-Russell diagram's main sequence depends on its mass. Small, cold
red dwarfs fuse slowly and stay on the main sequence for billions of years, while hot O-
type stars leave quickly after a few million years.
Mature stars
Eventually the star's core exhausts its supply of hydrogen, and the star begins to evolve
off the main sequence. Without the outward radiation pressure generated by the fusion
of hydrogen to counteract the force of gravity the core contracts until either electron
degeneracy pressure becomes sufficient to oppose gravity or the core becomes hot
enough (around 100 MK) for helium fusion to begin.
Low-mass stars
After low-mass stars cease fusion, their future is not directly observed due to the
universe's age. Models suggest 0.1 M☉ red dwarfs stay on the main sequence for
trillions of years, slowly evolving into white dwarfs. Slightly more massive stars become
red giants but don't reach the temperatures for helium fusion, transitioning to white
dwarfs after hydrogen shell burning. Mid-sized stars, around 0.6 M☉, fuse helium and
continue evolving beyond the red-giant branch.
Mid-sized stars
Stars of 0.6–10 M☉ become red giants, seen on the right of the Hertzsprung–Russell
diagram. Mid-sized stars experience two red giant phases: red-giant-branch with helium
cores and hydrogen-burning shells, and asymptotic-giant-branch with carbon cores and
helium-burning shells inside hydrogen-burning shells. They also spend time on the
horizontal branch with a helium-fusing core. Examples include Aldebaran and Arcturus.
Internal structures of main-sequence stars, convection zones with arrowed cycles
and radiative zones with red flashes. To the left a low-mass red dwarf, in the
center a mid-sized yellow dwarf and at the right a massive blue-white main-
sequence star.
Subgiant phase
When a star exhausts the hydrogen in its core, it leaves the main sequence and begins
to fuse hydrogen in a shell outside the core. The core increases in mass as the shell
produces more helium. Depending on the mass of the helium core, this continues for
several million to one or two billion years, with the star expanding and cooling at a
similar or slightly lower luminosity to its main sequence state. Eventually either the core
becomes degenerate, in stars around the mass of the sun, or the outer layers cool
sufficiently to become opaque, in more massive stars.
Red-giant-branch phase
As a star's outer layers become convective during the red-giant phase, fusion products
from deeper regions become visible at the surface for the first time. This stage reveals
subtle effects, such as changes in hydrogen and helium isotopes, detectable with
spectroscopy. The helium core grows, no longer in thermal equilibrium, causing
increased fusion in the hydrogen shell and a rise in luminosity toward the tip of the red-
giant branch.
Horizontal branch
In 0.6 to 2.0 solar mass stars, helium fusion ignites rapidly in a helium flash, releasing
immense energy. The core expands, slowing hydrogen fusion, and the star contracts,
migrating to the horizontal branch. Core helium flash stars evolve to the red end, while
higher-mass stars move along, with some becoming unstable pulsating stars or forming
a blue tail on the horizontal branch.
Asymptotic-giant-branch phase
After helium is depleted at the core, hydrogen and helium fusion continues in shells
around a carbon-oxygen core, following the asymptotic giant branch on the
Hertzsprung–Russell diagram with faster energy generation, marked by periodic thermal
pulses. During this phase, carbon stars can form through dredge-ups, and hot bottom
burning may convert carbon into oxygen and nitrogen, influencing luminosities and
spectra. Mira variables, another class of asymptotic giant branch stars, pulsate with
well-defined periods, and more massive stars exhibit longer periods and enhanced
mass loss, often becoming heavily obscured.
Post-AGB
Mid-range stars, reaching the tip of the asymptotic giant branch, undergo a post-
asymptotic-giant-branch super wind, producing a planetary nebula with a hot central
star that later cools to a white dwarf, expelling heavy-element-rich gas forming a
circumstellar envelope conducive to maser excitation.
The Cat's Eye Nebula, a planetary nebula formed by the death of a star with about the
same mass as the Sun.
It is possible for thermal pulses to be produced once post-asymptotic-giant-branch
evolution has begun, producing a variety of unusual and poorly understood stars known
as born-again asymptotic-giant-branch stars. These may result in extreme horizontal-
branch stars (subdwarf B stars), hydrogen deficient post-asymptotic-giant-branch stars,
variable planetary nebula central stars, and R Coronae Borealis variables.
Massive stars
In massive stars, the core is already large enough at the onset of the hydrogen burning
shell that helium ignition will occur before electron degeneracy pressure has a chance
to become prevalent. Thus, when these stars expand and cool, they do not brighten as
dramatically as lower-mass stars; however, they were more luminous on the main
sequence, and they evolve to highly luminous super giants.
Supergiant evolution
Extremely massive stars (over 40 M☉) retain high surface temperatures, preventing
them from becoming red super giants due to rapid mass loss from stellar winds. In
binary systems or rapidly rotating stars, lower-mass stars can also avoid becoming red
giants. The core of a massive star, after
helium fusion, grows hotter, reaching
temperatures sufficient for carbon and
heavier element fusion,
with stars over 8 M☉
producing neon, sodium,
and magnesium, while
slightly less massive
stars partially ignite
carbon before electron
degeneracy sets in.
The onion-like layers of a massive, evolved star just
before core collapse (not to scale)
For stars of approximately 8-12 M☉, after carbon burning, unstable neon burning
triggers a runaway fusion, resulting in an electron capture supernova; in more massive
stars, neon fusion proceeds without a runaway deflagration, followed by complete
oxygen and silicon burning, forming a core of iron-peak elements, which, upon reaching
an effective Chandrasekhar mass, leads to an electron capture-induced core collapse
and a supernova explosion.
Supernova
When the core of a massive star collapses, it forms either a neutron star or, if exceeding
the Tolman–Oppenheimer–
Volkoff limit, a black hole, and this
collapse releases gravitational
potential energy into a Type Ib,
Type Ic, or Type II supernova,
where the energetic neutrinos
produced enhance the shock
wave initiated by the rebounding
matter from the core collapse, leading to the creation of heavier-than-iron elements,
including radioactive ones up to uranium, a process distinct from the non-exploding red
giants' production of similar elements.
The Crab Nebula, the shattered remnants of a star which exploded as a
supernova visible in 1054 AD.
The energy transferred from the collapse of a massive star's core to rebounding
material generates heavy elements and accelerates them beyond escape velocity,
causing a Type Ib, Type Ic, or Type II supernova; while current computer models
partially account for the energy transfer, the observed ejection of material is not fully
explained, with neutrino oscillations and general-relativistic effects being potential
factors; pair-instability supernovae, occurring in the most massive stars, can completely
destroy the star without leaving a black hole remnants.
Stellar remnants
After a star has burned out its fuel supply, its remnants can take one of three forms,
depending on the mass during its lifetime.
White and black dwarfs
For a star of 1 M☉, the resulting white dwarf is of about 0.6 M☉, compressed into
approximately the volume of the Earth. White dwarfs are stable because the inward pull
of gravity is balanced by the degeneracy pressure of the star's electrons, a
consequence of the Pauli Exclusion Principle. Electron degeneracy pressure provides a
rather soft limit against further compression; therefore, for a given chemical
composition, white dwarfs of higher mass have a smaller volume. With no fuel left to
burn, the star radiates its remaining heat into space for billions of years.
This will lead either to collapse into a neutron star or runaway ignition of carbon and
oxygen. Heavier elements favor continued core collapse, because they require a higher
temperature to ignite, because electron capture onto these elements and their fusion
products is easier; higher core temperatures favor runaway nuclear reaction, which
halts core collapse and leads to a Type Ia supernova. These supernovae may be many
times brighter than the Type II supernova marking the death of a massive star, even
though the latter has the greater total energy release. This instability to collapse means
that no white dwarf more massive than approximately 1.4 M☉ can exist (with a possible
minor exception for very rapidly spinning white dwarfs, whose centrifugal force due to
rotation partially counteracts the weight of their matter). Mass transfer in a binary
system may cause an initially stable white dwarf to surpass the Chandrasekhar limit.
Neutron stars
Ordinarily, atoms are
mostly electron clouds by
volume, with very compact
nuclei at the center
(proportionally, if atoms
were the size of a football stadium, their nuclei would be the size of dust mites). When a
stellar core collapses, the pressure causes electrons and protons to fuse by electron
capture. Without electrons, which keep nuclei apart, the neutrons collapse into a dense
ball (in some ways like a giant atomic nucleus), with a thin overlying layer of degenerate
matter (chiefly iron unless matter of different composition is added later). The neutrons
resist further compression by the Pauli exclusion principle, in a way analogous to
electron degeneracy pressure, but stronger.
Bubble-like shock wave still expanding from a supernova explosion 15,000 years
ago
These stars, known as neutron stars, are extremely small—on the order of radius 10
km, no bigger than the size of a large city—and are phenomenally dense. Their period
of rotation shortens dramatically as the stars shrink (due to conservation of angular
momentum); observed rotational periods of neutron stars range from about 1.5
milliseconds (over 600 revolutions per second) to several seconds. When these rapidly
rotating stars' magnetic poles are aligned with the Earth, we detect a pulse of radiation
each revolution. Such neutron stars are called pulsars and were the first neutron stars
to be discovered. Though electromagnetic radiation detected from pulsars is most often
in the form of radio waves, pulsars have also been detected at visible, X-ray, and
gamma ray wavelengths.
Black holes
If the mass of the stellar remnant is high enough, the neutron degeneracy pressure will
be insufficient to prevent collapse below the Schwarzschild radius. The stellar remnant
thus becomes a black hole. The mass at which this occurs is not known with certainty
but is currently estimated at between 2 and 3 M☉.
Black holes are predicted by the theory of general relativity. According to classical
general relativity, no matter or information can flow from the interior of a black hole to an
outside observer, although quantum effects may allow deviations from this strict rule.
The existence of black holes in the universe is well supported, both theoretically and by
astronomical observation.
Because the core-collapse mechanism of a supernova is, at present, only partially
understood, it is still not known whether it is possible for a star to collapse directly to a
black hole without producing a visible supernova, or whether some supernovae initially
form unstable neutron stars which then collapse into black holes; the exact relation
between the initial mass of the star and the final remnant is also not completely certain.
Resolution of these uncertainties requires the analysis of more supernovae and
supernova remnants.
Models
A stellar evolutionary model is a mathematical model that can be used to compute the
evolutionary phases of a star from its formation until it becomes a remnant. The mass
and chemical composition of the star are used as the inputs, and the luminosity and
surface temperature are the only constraints. The model formulae are based upon the
physical understanding of the star, usually under the assumption of hydrostatic
equilibrium. Extensive computer calculations are then run to determine the changing
state of the star over time, yielding a table of data that can be used to determine the
evolutionary track of the star across the Hertzsprung–Russell diagram, along with other
evolving properties. Accurate models can be used to estimate the current age of a star
by comparing its physical properties with those of stars along a matching evolutionary
track.