Você está na página 1de 14

A red giant star is a dying star in the last stages of stellar evolution.

In only a few billion years, our


own sun will turn into a red giant star, expand and engulf the inner planets, possibly even Earth.
What does the future hold for the light of our solar system and others like it?
Forming a giant
Most of the stars in the universe are main sequence stars those converting hydrogen into
helium via nuclear fusion. A main sequence star may have a mass between a third to eight times
that of the sun and eventually burn through the hydrogen in its core. Over its life, the outward
pressure of fusion has balanced against the inward pressure of gravity. Once the fusion stops,
gravity takes the lead and compresses the star smaller and tighter.
Temperatures increase with the contraction, eventually reaching levels where helium is able to
fuse into carbon. Depending on the mass of the star, the helium burning might be gradual or might
begin with an explosive flash. The energy produced by the helium fusion causes the star to expand
outward to many times its original size.
Red giant stars reach sizes of 100 million to 1 billion kilometers in diameter (62 million to 621
million miles), 100 to 1,000 times the size of the sun today. Because the energy is spread across a
larger area, surface temperatures are actually cooler, reaching only 2,200 to 3,200 degrees Celsius
(4,000 to 5,800 degrees Fahrenheit), a little over half as hot as the sun. This temperature change
causes stars to shine in the redder part of the spectrum, leading to the name red giant, though
they are often more orangish in appearance.
Stars spend approximately a few thousand to 1 billion years as a red giant. Eventually, the helium
in the core runs out and fusion stops. The star shrinks again until a new helium shell reaches the
core. When the helium ignites, the outer layers of the star are blown off in huge clouds of gas and
dust known as planetary nebulae.
The core continues to collapse in on itself. Smaller stars such as the sun end their lives as compact
white dwarfs. The material of larger, more massive stars fall inward until the star eventually
becomes a supernova, blowing off gas and dust in a dramatic fiery death
he CNO cycle (for carbonnitrogenoxygen) is one of the two known sets of fusion reactions by
which stars convert hydrogen to helium, the other being the protonproton chain reaction. Unlike
the latter, the CNO cycle is a catalytic cycle. It is dominant in stars that are more than 1.3 times
as massive as the Sun.[1]
In the CNO cycle, four protons fuse, using carbon, nitrogen and oxygen isotopes as catalysts, to
produce one alpha particle, two positrons and two electron neutrinos. Although there are various
paths and catalysts involved in the CNO cycles, all these cycles have the same net result:
41
1H
+ 2
e
4
2He
+ 2
e+
+ 2
e
+ 2

e + 3

+ 24.7 MeV 4
2He
+ 2

e + 3

+ 26.7 MeV
The positrons will almost instantly annihilate with electrons, releasing energy in the form
of gamma rays. The neutrinos escape from the star carrying away some energy. One nucleus goes
to become carbon, nitrogen, and oxygen isotopes through a number of transformations in an
endless loop.
Theoretical models suggest that the CNO cycle is the dominant source of energy in stars whose
mass is greater than about 1.3 times that of the Sun.[1] The protonproton chain is more important
in stars the mass of the Sun or less. This difference stems from temperature dependency
differences between the two reactions; pp-chain reaction starts at temperatures
around 4106 K[2] (4 megakelvins), making it the dominant energy source in smaller stars. A self-
maintaining CNO chain starts at approximately 15106 K, but its energy output rises much more
rapidly with increasing temperatures.[1] At approximately 17106 K, the CNO cycle starts becoming
the dominant source of energy.[3] The Sun has a core temperature of around 15.7106 K, and
only 1.7% of 4
He
nuclei produced in the Sun is born in the CNO cycle. The CNO-I process was independently
proposed by Carl von Weizscker[4] and Hans Bethe[5] in 1938 and 1939, respectively.
Cold CNO cycles[edit]
Under typical conditions found in stars, catalytic hydrogen burning by the CNO cycles is limited by
proton captures. Specifically, the timescale for beta decay of the radioactive nuclei produced is
faster than the timescale for fusion. Because of the long timescales involved, the cold CNO cycles
convert hydrogen to helium slowly, allowing them to power stars in quiescent equilibrium for
many years.
CNO-I[edit]
The first proposed catalytic cycle for the conversion of hydrogen into helium was initially called the
carbonnitrogen cycle (CN cycle), also honorarily referred to as the BetheWeizscker cycle,
because it does not involve a stable isotope of oxygen. Bethe's original calculations suggested the
CN-cycle was the Sun's primary source of energy, owing to the belief at the time that the Sun's
composition was 10% nitrogen;[5] the solar abundance of nitrogen is now known to be less than
half a percent. This cycle is now recognized as the first part of the larger CNO nuclear burning
network. The main reactions of the CNO-I cycle are 12
6C
13
7N
13
6C
14
7N
15
8O
15
7N
12
6C
:[6]
12 1 13
1.95 Me
6C + 1H 7N + +
V

13 13
7N 6C + e+ + + 1.20 MeV (half-life of 9.965 minutes[7])
e
13 1 14
6C + 1H 7N + + 7.54 MeV

14 1 15
7N + 1H 8O + + 7.35 MeV

15 15
8O 7N + e+ + + 1.73 MeV (half-life of 122.24 seconds[7])
e
15 1 12 4
7N + 1H 6C + 2He + 4.96 MeV

where the carbon-12 nucleus used in the first reaction is regenerated in the last reaction. After the
two positrons emitted annihilate with two ambient electrons producing an additional 2.04 MeV,
the total energy released in one cycle is 26.73 MeV; it should be noted that in some texts, authors
are erroneously including the positron annihilation energy in with the beta-decay Q-value and
then neglecting the equal amount of energy released by annihilation, leading to possible
confusion. All values are calculated with reference to the Atomic Mass Evaluation 2003. [8]
The limiting (slowest) reaction in the CNO-I cycle is the proton capture on 14
7N
. In 2006 it was experimentally measured down to stellar energies, revising the calculated age
of globular clusters by around 1 billion years.[9]
The neutrinos emitted in beta decay will have a spectrum of energy ranges, because
although momentum is conserved, the momentum can be shared in any way between the
positron and neutrino, with either emitted at rest and the other taking away the full energy, or
anything in between, so long as all the energy from the Q-value is used. All momentum which get
the electron and the neutrino together is not great enough to cause a significant recoil of the
much heavier daughter nucleus and hence, its contribution to kinetic energy of the products, for
the precision of values given here, can be neglected. Thus the neutrino emitted during the decay
of nitrogen-13 can have an energy from zero up to 1.20 MeV, and the neutrino emitted during the
decay of oxygen-15 can have an energy from zero up to 1.73 MeV. On average, about 1.7 MeV of
the total energy output is taken away by neutrinos for each loop of the cycle, leaving about 25
MeV available for producing luminosity.[10]
Use in astronomy[edit]
While the total number of "catalytic" nuclei are conserved in the cycle, in stellar evolution the
relative proportions of the nuclei are altered. When the cycle is run to equilibrium, the ratio of the
carbon-12/carbon-13 nuclei is driven to 3.5, and nitrogen-14 becomes the most numerous
nucleus, regardless of initial composition. During a star's evolution, convective mixing episodes
moves material, within which the CNO cycle has operated, from the star's interior to the surface,
altering the observed composition of the star. Red giant stars are observed to have lower carbon-
12/carbon-13 and carbon-12/nitrogen-14 ratios than do main sequence stars, which is considered
to be convincing evidence for the operation of the CNO cycle
Supernova explosions
Supernova
Iron cannot release energy by fusion because it requires a larger input of energy than it releases.
So the iron core continues to be subjected to gravity, which pushes the electrons closer to the
nuclei than the quantum limit allows, and they disappear by combining with protons to form
neutrons, giving off neutrinos in the process. Once this process starts, in a fraction of a second, an
iron core the size of the earth and with a mass like our Sun, collapses into a ball of neutrons a few
kilometers across. This gravitational collapse releases an enormous amount of energy, more than
100 times what our Sun will radiate over its entire 10 billion year lifetime. This energy blows the
outer layers of the star off into space in a giant explosion called a supernova (plural: supernovae.)
The ball of neutrons left behind is called a neutron star and is incredibly dense. In some cases the
remaining mass is large enough that gravity continues to collapse the core until it becomes a black
hole.
The explosion sends a shock wave of the
star's former surface zooming out at a speed of 10,000 km/s, and heating it so it shines brilliantly
for about a week. This shock wave compresses the material it passes through and is the only place
where many elements such as zinc, silver, tin, gold, mercury, lead and uranium are produced. Over
several months the gases cool and fade in brightness and join the debris of interstellar space. This
debris has in it all of the elements that were created in the star's core. Millions or billions of years
later, this debris may be incorporated into new stars. The fact that the Earth contains elements
that are produced only in supernovae is evidence that our solar system, planet and bodies contain
material that was produced long ago by a supernova.
The crab nebula is a remnant from a supernova that went off in 1054 A.D. When Betelgeuse
explodes as a supernova it will be more than 10 times brighter than the full moon in our sky. It is
only 640 light years away, and could have already become a supernova, but the light from it just
hasn't reached us yet.
Supernovae occur in stars with at least 8 solar masses.
The rapid neutron capture process or r-process is a set of reactions in nuclear astrophysics that
are responsible for the creation (nucleosynthesis) of approximately half the atomic nuclei heavier
than iron.
This process entails a succession of rapid neutron captures (hence the name) by heavy seed nuclei,
typically beginning with 56Fe. The captures must be rapid in the sense that the nucleus does not
have time to undergo radioactive decay before another neutron arrives to be captured. The r-
process therefore occurs in locations where there is a high flux of free neutrons. These include the
material ejected from a core-collapse supernovae (as part of supernova nucleosynthesis) and
neutron-rich matter thrown off from a neutron star merger. The relative contributions of these,
and other, sources to the astrophysical abundance of r-process elements is a matter of ongoing
research.
The r-process also occurs to a minor extent in thermonuclear weapon explosions. This led to the
historical discovery of the elements einsteinium (element 99) and fermium (element 100) in
nuclear weapon fallout.
The r-process is distinguished from the s-process, the other predominant mechanism for the
production of heavy elements, which is nucleosynthesis by means of slow captures of neutrons.
The s-process primarily occurs within stars, particularly AGB stars, where the neutron flux is
sufficient to cause reactions but too low for the r-process. The s-process is secondary, meaning
that it requires pre-existing heavy isotopes as seed nuclei to be converted into other heavy nuclei.
Taken together, the r- and s-processes account for the majority of abundance evolution of
elements heavier than iron.
History[edit]
The need for some kind of rapid capture of neutrons was seen from the relative abundances of
isotopes of heavy elements given in a newly published table of abundances by Hans
Suess and Harold Urey in 1956. Radioactive isotopes must capture another neutron faster than
they can undergo beta decay in order to create abundance peaks at germanium, xenon,
and platinum. According to the nuclear shell model, radioactive nuclei that would decay
into isotopes of these elements have closed neutron shells near the neutron drip line, where more
neutrons cannot be added. Those abundance peaks created by rapid neutron capture implied that
other nuclei could be accounted for by such a process. That process of rapid neutron capture in
neutron-rich isotopes is called the r-process. A table apportioning the heavy isotopes
phenomenologically between s-process and r-process was published in the famous B2FH review
paper in 1957,[1] which named that process and outlined the physics that guide it. B 2FH also
elaborated the theory of stellar nucleosynthesis and set substantial frame-work for
contemporary nuclear astrophysics.
The r-process described by the B2FH paper was first computed time-dependently at Caltech by
Phillip Seeger, William A. Fowler and Donald D. Clayton,[2] who achieved the first successful
characterization of the r-process abundances and showed its evolution in time. They were also
able using theoretical production calculations to construct more quantitative apportionment
between s-process and r-process of the abundance table of heavy isotopes, thereby establishing a
more reliable abundance curve for the r-process isotopes than B2FH had been able to define.
Today, the r-process abundances are determined using their technique of subtracting the more
reliable s-process isotopic abundances from the total isotopic abundances and attributing the
remainder to the r-process nucleosynthesis. That r-process abundance curve (vs. atomic weight)
gratifyingly resembles computations of abundances synthesized by the physical process.
Most neutron-rich isotopes of elements heavier than nickel are produced, either exclusively or in
part, by the beta decay of very radioactive matter synthesized during the r-process by rapid
absorption, one after another, of free neutrons created during the explosions. The creation of free
neutrons by electron capture during the rapid collapse to high density of the supernova core along
with assembly of some neutron-rich seed nuclei makes the r-process a primary process; namely,
one that can occur even in a star of pure H and He, in contrast to the B2FH designation as
a secondary process building on preexisting iron.
Observational evidence of the r-process enrichment of stars, as applied to the abundance
evolution of the galaxy of stars, was laid out by Truran in 1981.[3] He and many subsequent
astronomers showed that the pattern of heavy-element abundances in the earliest metal-poor
stars matched that of the shape of the solar r-process curve, as if the s-process component were
missing. This was consistent with the hypothesis that the s-process had not yet begun in these
young stars, for it requires about 100 million years of galactic history to get started. These stars
were born earlier than that, showing that the r-process emerges immediately from quickly-
evolving massive stars that become supernovae. The primary nature of the r-process from
observed abundance spectra in old stars born when the galactic metallicity was still small but that
nonetheless contain their complement of r-process nuclei.

Periodic table showing the cosmogenic origin of each element. The elements heavier than iron
with origins in supernovae are typically those produced by the r-process, which is powered by
supernovae neutron bursts
This scenario, though generally supported by supernova experts, has yet to achieve a totally
satisfactory calculation of r-process abundances because the overall problem is numerically
formidable; but existing results are very supportive.
The r-process is responsible for our natural cohort of radioactive elements, such as uranium and
thorium, as well as the most neutron-rich isotopes of each heavy element.
Astrophysical sites[edit]
The most probable candidate sites for the r-process has long been suggested to be core-
collapse supernovae (spectral Type Ib, Ic and II), which may provide the necessary physical
conditions for the r-process. However, the abundance of r-process nuclei requires that either only
a small fraction of supernovae eject r-process nuclei to the interstellar medium, or that each
supernova ejects only a very small amount of r-process material. In addition, the ejected material
must be relatively neutron-rich, a condition which has been difficult to achieve in models.[7] An
alternative site proposed in 1974[8] was decompressing neutron star matter. It was proposed such
matter is ejected from neutron stars merging with black holes in compact binaries. In 1989[9] (see
also [10]) this scenario was extended to binary neutron star mergers (a binary star system of two
neutron stars that collide). These sites may now be starting to be observationally confirmed.[

Stellar nucleosynthesis is the process by which the natural abundances of the chemical
elements within stars change due to nuclear fusion reactions in the cores and their overlying
mantles. Stars are said to evolve (age) with changes in the abundances of the elements within.
Core fusion increases the atomic weight of elements and reduces the number of particles, which
would lead to a pressure loss except that gravitation leads to contraction, an increase of
temperature, and a balance of forces.[1] A star loses most of its mass when it is ejected late in the
star's stellar lifetimes, thereby increasing the abundance of elements heavier than helium in
the interstellar medium. The term supernova nucleosynthesis is used to describe the creation of
elements during the evolution and explosion of a presupernova star, a concept put forth by Fred
Hoyle in 1954.[2] A stimulus to the development of the theory of nucleosynthesis was the discovery
of variations in the abundances of elements found in the universe. Those abundances, when
plotted on a graph as a function of atomic number of the element, have a jagged sawtooth shape
that varies by factors of tens of millions. This suggested a natural process other than random. Such
a graph of the abundances can be seen at History of nucleosynthesis theory article. Of the several
processes of nucleosynthesis, stellar nucleosynthesis is the dominating contributor to elemental
abundances in the universe.
A second stimulus to understanding the processes of stellar nucleosynthesis occurred during the
20th century, when it was realized that the energy released from nuclear fusion reactions
accounted for the longevity of the Sun as a source of heat and light.[3] The fusion of nuclei in a star,
starting from its initial hydrogen and helium abundance, provides it energy and the synthesis of
new nuclei is a byproduct of that fusion process. This became clear during the decade prior
to World War II. The fusion-produced nuclei are restricted to those only slightly heavier than the
fusing nuclei; thus they do not contribute heavily to the natural abundances of the elements.
Nonetheless, this insight raised the plausibility of explaining all of the natural abundances of
elements in this way. The prime energy producer in our Sun is the fusion of hydrogen to
form helium, which occurs at a solar-core temperature of 14 million kelvin.
History[edit]

In 1920 Arthur Eddington proposed that stars obtained their energy from nuclear
fusion of hydrogen to form helium and raised the possibility that the heavier elements are
produced in stars.
In 1920, Arthur Eddington, on the basis of the precise measurements of atomic masses by F.W.
Aston and a preliminary suggestion by Jean Perrin, proposed that stars obtained their energy
from nuclear fusion of hydrogen to form helium and raised the possibility that the heavier
elements are produced in stars.[4][5][6] This was a preliminary step toward the idea of
nucleosynthesis. In 1928, George Gamow derived what is now called the Gamow factor,
a quantum-mechanical formula that gave the probability of bringing two nuclei sufficiently close
for the strong nuclear force to overcome the Coulomb barrier. The Gamow factor was used in the
decade that followed by Atkinson and Houtermans and later by Gamow himself and Edward
Teller to derive the rate at which nuclear reactions would proceed at the high temperatures
believed to exist in stellar interiors.
In 1939, in a paper entitled "Energy Production in Stars", Hans Bethe analyzed the different
possibilities for reactions by which hydrogen is fused into helium.[7] He defined two processes that
he believed to be the sources of energy in stars. The first one, the protonproton chain reaction, is
the dominant energy source in stars with masses up to about the mass of the Sun. The second
process, the carbonnitrogenoxygen cycle, which was also considered by Carl Friedrich von
Weizscker in 1938, is most important in more massive stars. These works concerned the energy
generation capable of keeping stars hot. A clear physical description of the protonproton chain
and of the CNO cycle appears in a 1968 textbook.[8] Bethe's two papers did not address the
creation of heavier nuclei, however. That theory was begun by Fred Hoyle in 1946 with his
argument that a collection of very hot nuclei would assemble into iron.[9] Hoyle followed that in
1954 with a large paper describing how advanced fusion stages within stars would synthesize
elements between carbon and iron in mass.[10] This is the dominant work in stellar
nucleosynthesis.[11] It provided the roadmap to how the most abundant elements on Earth had
been synthesized from initial hydrogen and helium, making clear how those abundant elements
increased their galactic abundances as the galaxy aged.
Quickly, Hoyle's theory was expanded to other processes, beginning with the publication of a
celebrated review paper in 1957 by Burbidge, Burbidge, Fowler and Hoyle (commonly referred to
as the B2FH paper).[12] This review paper collected and refined earlier research into a heavily cited
picture that gave promise of accounting for the observed relative abundances of the elements; but
it did not itself enlarge Hoyle's 1954 picture for the origin of primary nuclei as much as many
assumed, except in the understanding of nucleosynthesis of those elements heavier than iron.
Significant improvements were made by Alastair GW Cameron and by Donald D. Clayton. Cameron
presented his own independent approach[13] (following Hoyle's approach for the most part) of
nucleosynthesis. He introduced computers into time-dependent calculations of evolution of
nuclear systems. Clayton calculated the first time-dependent models of the S-process[14] and of
the R-process,[15] as well as of the burning of silicon into the abundant alpha-particle nuclei and
iron-group elements,[16] and discovered radiogenic chronologies[17] for determining the age of the
elements. The entire research field expanded rapidly in the 1970s.
The triple-alpha process is a set of nuclear fusion reactions by which three helium-4 nuclei (
See alpha process for more details about this reaction and further steps in the chain of stellar
nucleosynthesis.
This creates a situation in which stellar nucleosynthesis produces large amounts of carbon and
oxygen but only a small fraction of those elements are converted into neon and heavier elements.
Both oxygen and carbon make up the 'ash' of helium-4 burning.
cles) are transformed into carbon.[
Reaction rate and stellar evolution[edit]
The triple-alpha steps are strongly dependent on the temperature and density of the stellar
material. The power released by the reaction is approximately proportional to the temperature to
the 40th power, and the density squared.[6] In contrast, the PP chain produces energy at a rate
proportional to the fourth power of temperature, the CNO cycle at about the 17th power of the
temperature, and both are linearly proportional to the density. This strong temperature
dependence has consequences for the late stage of stellar evolution, the red giant stage.
For lower mass stars, the helium accumulating in the core is prevented from further collapse only
by electron degeneracy pressure. As the temperature rises, increased pressure in the core would
normally result in an expansion, reduction of density, and thus reduction in reaction rate.
However, due to the high pressure at the center of the star this does not occur and energy
production continues unmoderated. As a consequence, the temperature increases, causing the
reaction rate further increase in a positive feedback cycle that becomes a runaway reaction. This
process, known as the helium flash, lasts a matter of seconds but burns 6080% of the helium in
the core. During the core flash, the star's energy production can reach approximately 1011 solar
luminosities which is comparable to the luminosity of a whole galaxy,[7] although no effects will be
immediately observed at the surface, as it is hidden by the star's overlying layers.
For higher mass stars, carbon collects in the core, displacing the helium to a surrounding shell
where helium burning occurs. In this helium shell, the pressures are lower and the mass is not
supported by electron degeneracy. Thus, as opposed to the center of the star, the shell is able to
expand in response to increased thermal pressure in the helium shell. Expansion cools this layer
and slows the reaction, causing the star to contract again. This process continues cyclically, and
stars undergoing this process will have periodically variable radius and power production. These
stars will also lose material from their outer layers as they expand and contract.
Discovery[edit]
The triple alpha process is highly dependent on carbon-12 and beryllium-8 having resonances with
slightly more energy than helium-4, and before 1952, no such energy levels were known for
carbon. The astrophysicist Fred Hoyle used the fact that carbon-12 is abundant in the universe as
evidence for the existence of a carbon-12 resonance. The only way Hoyle could find that would
produce an abundance of both carbon and oxygen is through a triple alpha process with a carbon-
12 resonance near 7.68 MeV.[8]
Hoyle went to nuclear physicist William Alfred Fowler's lab at Caltech and said that there had to be
a resonance of 7.68 MeV in the carbon-12 nucleus. (There had been reports of an excited state at
about 7.5 MeV.[8]) Fred Hoyle's audacity in doing this is remarkable, and initially the nuclear
physicists in the lab were skeptical. Finally, a junior physicist, Ward Whaling, fresh from Rice
University, who was looking for a project decided to look for the resonance. Fowler gave Whaling
permission to use an old Van de Graaff generator that was not being used. Hoyle was back in
Cambridge when his prediction was verified a few months later. The nuclear physicists put Hoyle
as first author on a paper delivered by Whaling at the Summer meeting of the American Physical
Society. A long and fruitful collaboration between Hoyle and Fowler soon followed, with Fowler
even coming to Cambridge.[9] By 1952, Fowler had noted the beryllium-8 resonance, and Edwin
Salpeter calculated the reaction rate taking this resonance into account.[10][11]
This helped to explain the rate of the process, but the rate calculated by Salpeter seemed too low
at the temperatures expected in supernovas.[8] When Fowler's lab discovered a carbon-12
resonance near 7.65 MeV it eliminated the discrepancy between the nuclear theory and the
theory of stellar evolution.
The final reaction product lies in a 0+ state (0 spin and positive parity). Since the Hoyle state was
predicted to be either a 0+ or a 2+ state, electronpositron pairs or gamma rays were expected to
be seen. However, when experiments were carried out, the gamma emission reaction channel was
not observed, and this meant the state must be a 0+ state. This state completely suppresses single
gamma emission, since single gamma emission must carry away at least 1 unit of angular
momentum. Pair production from an excited 0+ state is possible because their combined spins (0)
can couple to a reaction that has a change in angular momentum of 0.[12]
Supernova nucleosynthesis is a theory of the production of many different chemical
elements in supernova explosions, first advanced by Fred Hoyle in 1954.[1] The nucleosynthesis, or
fusion of lighter elements into heavier ones, occurs during explosive oxygen burning and silicon
burning processes.[2] Those fusion reactions create the
elements silicon, sulfur, chlorine, argon, sodium, potassium, calcium, scandium, titanium and iron
peak elements: vanadium, chromium, manganese, iron, cobalt, and nickel. These are called
"primary elements", in that they can be fused from pure hydrogen and helium in massive stars. As
a result of their ejection from supernovae, their abundances increase within the interstellar
medium. Elements heavier than nickel are created primarily by a rapid capture of neutrons in a
process called the r-process. However, these are much less abundant than the primary chemical
elements. Other processes thought to be responsible for some of the nucleosynthesis of
underabundant heavy elements, notably a proton capture process known as the rp-process and
a photodisintegration process known as the gamma (or p) process. The latter synthesizes the
lightest, most neutron-poor, isotopes of the heavy elements. A supernova is a massive explosion
of a star that occurs under two principal scenarios. The first is that a white dwarf star undergoes a
nuclear-based explosion after it reaches its Chandrasekhar limit after absorbing mass from a
neighboring star (usually a red giant). The second, and more common, cause is when a massive
star, usually a supergiant, reaches nickel-56 in its nuclear fusion (or burning) processes. This
isotope undergoes radioactive decay into iron-56, which has one of the highest binding energies of
all of the isotopes, and is the last element that produces a net release of energy by nuclear
fusion, exothermically.
All nuclear fusion reactions that produce heavier elements cause the star to lose energy and are
said to be endothermic reactions. The pressure that supports the star's outer layers drops sharply.
As the outer envelope is no longer sufficiently supported by the radiation pressure, the star's
gravity pulls its outer layers rapidly inward. As the star collapses, these outer layers collide with
the incompressible stellar core, producing a shockwave that expands outward through the
unfused material of the outer shell. The pressures and densities in the shockwave are sufficient to
induce fusion in that material, and the energy released leads to the star's explosion, dispersing
material from the star into interstellar space
The slow neutron capture process or s-process is a series of reactions in nuclear
astrophysics which occur in stars, particularly AGB stars. The s-process is responsible for the
creation (nucleosynthesis) of approximately half the atomic nuclei heavier than iron.
In the s-process, a seed nucleus undergoes neutron capture to form an isotope with one
higher atomic mass. If the new isotope is stable a series of increases in mass can occur, but if it
is unstable then beta decay will occur, producing an element of the next highest atomic number.
The process is slow (hence the name) in the sense that there is sufficient time for this radioactive
decay to occur before another neutron is captured. A series of these reactions produces stable
isotopes by moving along the valley of beta-decay stable isobars in the chart of isotopes.
A range of elements and isotopes can be produced by the s-process, because of the intervention
of alpha decay steps along the reaction chain. The relative abundances of elements and isotopes
produced depends on the source of the neutrons and how their flux changes over time. Each
branch of the s-process reaction chain eventually terminates at a cycle involving lead, bismuth,
and polonium.
The s-process contrasts with the r-process, in which successive neutron captures are rapid: they
happen more quickly than the beta decay can occur. The r-process dominates in environments
which have a higher flux of free neutrons; it produces heavier elements and more neutron-rich
isotopes than the s-process. Together the two processes account for the majority of abundance
evolution of elements heavier than iron.
History[edit]
The s-process was seen to be needed from the relative abundances of isotopes of heavy elements
and from a newly published table of abundances by Hans Suess and Harold Urey in 1956. Among
other things, these data showed abundance peaks for strontium, barium, and lead, which,
according to quantum mechanics and the nuclear shell model, are particularly stable nuclei, much
like the noble gases are chemically inert. This implied that some abundant nuclei must be created
by slow neutron capture, and it was only a matter of determining how other nuclei could be
accounted for by such a process. A table apportioning the heavy isotopes between s-process and r-
process was published in the famous B2FH review paper in 1957.[1] There it was also argued that
the s-process occurs in red giant stars. In a particularly illustrative case, the element technetium,
whose longest half-life is 4.2 million years, had been discovered in S-, M-, and N-type stars in
1952.[2][3] Since these stars were thought to be billions of years old, the presence of technetium in
their outer atmospheres was taken as evidence of its recent creation there, probably unconnected
with the nuclear fusion in the deep interior of the star that provides its power.

Periodic table showing the cosmogenic origin of each element. The elements heavier than iron
with origins in large stars are typically those produced by the s-process, which is characterized by
slow neutron diffusion and capture over long periods in such stars
A calculable model for creating the heavy isotopes from iron seed nuclei in a time-dependent
manner was not provided until 1961.[4] That work showed that the large overabundances of
barium observed by astronomers in certain red-giant stars could be created from iron seed nuclei
if the total neutron flux (number of neutrons per unit area) was appropriate. It also showed that
no one single value for neutron flux could account for the observed s-process abundances, but
that a wide range is required. The numbers of iron seed nuclei that were exposed to a given flux
must decrease as the flux becomes stronger. This work also showed that the curve of the product
of neutron-capture cross section times abundance is not a smoothly falling curve, as B2FH had
sketched, but rather has a ledge-precipice structure. A series of papers[5][6][7][8][9] in the 1970s
by Donald D. Clayton utilizing an exponentially declining neutron flux as a function of the number
of iron seed exposed became the standard model of the s-process and remained so until the
details of AGB-star nucleosynthesis became advanced enough that they became a standard model
based on the stellar structure models. Important series of measurements of neutron-capture cross
sections were reported from Oak Ridge National Lab in 1965[10] and by Karlsruhe Nuclear Physics
Center in 1982[11] and subsequently. These placed the s-process on the firm quantitative basis that
it enjoys today.
The s-process in stars[edit]
The s-process is believed to occur mostly in asymptotic giant branch stars. In contrast to the r-
process which is believed to occur over time scales of seconds in explosive environments, the s-
process is believed to occur over time scales of thousands of years, passing decades between
neutron captures. The extent to which the s-process moves up the elements in the chart of
isotopes to higher mass numbers is essentially determined by the degree to which the star in
question is able to produce neutrons. The quantitative yield is also proportional to the amount of
iron in the star's initial abundance distribution. Iron is the "starting material" (or seed) for this
neutron capture beta-minus decay sequence of synthesizing new elements.
The main neutron source reactions are:

Você também pode gostar