Вы находитесь на странице: 1из 16

Chemical reactions

Chemical reactions and how they break and form bonds between atoms. Balanced reactions,
reversibility, and equilibrium.
Introduction
Molecules—like the ones that make up your body—are just collections of atoms held together by
chemical bonds. In many ways, they're a lot like Tinkertoy® building projects. In fact, if you take organic
chemistry, you’ll most likely buy a model set that looks suspiciously similar to Tinkertoys®:

Ball-and-stick model of the atom proline made using a modeling kit.


_Image credit: "Proline model," by Peter Murray-Rust (CC BY-SA 2.5._

Just as you can put Tinkertoy® wheels together in different ways using different stick connectors, you
can also put atoms together in a different ways by forming different sets of chemical bonds. The process
of reorganizing atoms by breaking one set of chemical bonds and forming a new set is known as a
chemical reaction.

Chemical reactions

Chemical reactions occur when chemical bonds between atoms are formed or broken. The substances
that go into a chemical reaction are called the reactants, and the substances produced at the end of
the reaction are known as the products. An arrow is drawn between the reactants and products to
indicate the direction of the chemical reaction, though a chemical reaction is not always a "one-way
street," as we'll explore further in the next section.

For example, the reaction for breakdown of hydrogen peroxide (\text{H}_{2}H2H, start subscript, 2, end
subscript\text{O}_{2}O2O, start subscript, 2, end subscript) into water and oxygen can be written as:

2 \text{H}_{2}2H22, H, start subscript, 2, end subscript\text{O}_{2} \text{(hydrogen peroxide)}O2


(hydrogen peroxide)O, start subscript, 2, end subscript, left parenthesis, h, y, d, r, o, g, e, n, space, p, e,
r, o, x, i, d, e, right parenthesis \rightarrow→right arrow 2\text{H}_{2}\text O \text{(water)}2H2
O(water)2, H, start subscript, 2, end subscript, O, left parenthesis, w, a, t, e, r, right
parenthesis + \text{O}_{2}\text{(oxygen)}O2(oxygen)O, start subscript, 2, end subscript, left
parenthesis, o, x, y, g, e, n, right parenthesis

In this example hydrogen peroxide is our reactant, and it gets broken down into water and oxygen, our
products. The atoms that started out in hydrogen peroxide molecules are rearranged to form water
molecules (\text{H}_{2}\text OH2OH, start subscript, 2, end subscript, O) and oxygen molecules (\text
O_2O2O, start subscript, 2, end subscript).

You may have noticed extra numbers in the chemical equation above: the 222s in front of hydrogen
peroxide and water. These numbers are called coefficients, and they tell us how many of each molecule
participate in the reaction. They must be included in order to make our equation balanced, meaning
that the number of atoms of each element is the same on the two sides of the equation.

Equations must be balanced to reflect the law of conservation of matter, which states that no atoms
are created or destroyed over the course of a normal chemical reaction. You can learn more about
balancing reactions in the balancing chemical equations tutorial.

Reversibility and equilibrium

Some chemical reactions simply run in one direction until the reactants are used up. These reactions
are said to be irreversible. Other reactions, however, are classified as reversible. Reversible
reactions can go in both the forward and backward directions.

In a reversible reaction, reactants turn into products, but products also turn back into reactants. In fact,
both the forward reaction and its opposite will take place at the same time. This back and forth
continues until a certain relative balance between reactants and products is reached—a state
called equilibrium. At equilibrium, the forward and backward reactions are still happening, but the
relative concentrations of products and reactants no longer change.

Each reaction has its own characteristic equilibrium point, which we can describe with a number called
the equilibrium constant. To learn where the equilibrium constant comes from and how to calculate it
for a specific reaction, check out the equilibrium topic.

When a reaction is classified as reversible, it is usually written with paired forward and backward arrows
to show it can go both ways. For example, in human blood, excess hydrogen ions (\text H^+H+H, start
superscript, plus, end superscript) bind to bicarbonate ions (\text{HCO}_{3}HCO3H, C, O, start subscript,
3, end subscript^{-}−start superscript, minus, end superscript), forming carbonic acid (\text{H}_{2}H2H,
start subscript, 2, end subscript\text{CO}_{3}CO3C, O, start subscript, 3, end subscript):

\text{HCO}_{3}HCO3H, C, O, start subscript, 3, end subscript^{-}−start superscript, minus, end


superscript + \text{H}^{+}H+H, start superscript, plus, end
superscript \rightleftharpoons⇌ \text{H}_{2}H2H, start subscript, 2, end subscript\text{CO}_{3}CO3C,
O, start subscript, 3, end subscript

Since this is a reversible reaction, if carbonic acid were added to the system, some of it would be turned
into bicarbonate and hydrogen ions to restore equilibrium. In fact, this buffer system plays a key role
in keeping your blood pH stable and healthy.
How are light and heavy elements formed? (Advanced)

For an independent study course at my high school, I am researching the formation of the elements,
both light and heavy. I have some basic understanding of how this is done, and I have also found some
technical information that at this time I don't understand. Can you point me to some good articles on
the topic, or perhaps cover some more advanced materials yourself?

The lightest elements (hydrogen, helium, deuterium, lithium) were produced in the Big Bang
nucleosynthesis. According to the Big Bang theory, the temperatures in the early universe were so high
that fusion reactions could take place. This resulted in the formation of light elements: hydrogen,
deuterium, helium (two isotopes), lithium and trace amounts of beryllium.

Nuclear fusion in stars converts hydrogen into helium in all stars. In stars less massive than the Sun, this
is the only reaction that takes place. In stars more massive than the Sun (but less massive than about 8
solar masses), further reactions that convert helium to carbon and oxygen take place in succesive stages
of stellar evolution. In the very massive stars, the reaction chain continues to produce elements like
silicon up to iron.

Elements higher than iron cannot be formed through fusion as one has to supply energy for the reaction
to take place. However, we do see elements higher than iron around us. So how did these elements
form? The answer is supernovae. In a supernova explosion, neutron capture reactions take place (this
is not fusion), leading to the formation of heavy elements. This is the reason why it is said that most of
the stuff that we see around us come from stars and supernovae (the heavy elements part). If you go
into technical details, then there are two processes of neutron capture called rapid process (r-process)
and the slow process (s-process), and these lead to formation of different elements.

The proton–proton chain reaction is one of two known sets of nuclear fusion reactions by
which stars convert hydrogen to helium. It dominates in stars with masses less than or equal to that of
the Sun,[1] whereas the CNO cycle, the other known reaction, is suggested by theoretical models to
dominate in stars with masses greater than about 1.3 times that of the Sun.[2]
In general, proton–proton fusion can occur only if the kinetic energy (i.e. temperature) of the protons
is high enough to overcome their mutual electrostatic or Coulomb repulsion.[3]
In the Sun, deuterium-producing events are rare. Diprotons are the much more common result of
proton–proton reactions within the star, and diprotons almost immediately decay back into two
protons. Since the conversion of hydrogen to helium is slow, the complete conversion of the hydrogen
in the core of the Sun is calculated to take more than ten billion years.[4]
Although often called the "proton–proton chain reaction", it is not a chain reaction in the normal sense
of the word (at least not branch I – in branches II and III, helium, which is the product, also serves as a
catalyst). It does not produce particles that go on to induce the reaction to continue (such as neutrons
given off during fission). In fact, the rate is self-limiting because the heat produced tends toward
reducing the density. It is however a chain (like a decay chain) and a reaction, or more accurately a
branched chain of reactions starting with two protons coming together and yielding deuterium.
CNO CYCLE
The ‘CNO cycle’ refers to the Carbon-Nitrogen-Oxygen cycle, a process of stellar nucleosynthesis in
which stars on the Main Sequence fuse hydrogen into helium via a six-stage sequence of reactions.
This sequence proceeds as follows:

 A carbon-12 nucleus captures a proton and emits a gamma ray, producing nitrogen-13.
 Nitrogen-13 is unstable and emits a beta particle, decaying to carbon-13.
 Carbon-13 captures a proton and becomes nitrogen-14 via emission of a gamma-ray.
 Nitrogen-14 captures another proton and becomes oxygen-15 by emitting a gamma-ray.
 Oxygen-15 becomes nitrogen-15 via beta decay.
 Nitrogen-15 captures a proton and produces a helium nucleus (alpha particle) and carbon-12, which
is where the cycle started.

Thus, the carbon-12 nucleus used in the initial reaction is regenerated in the final one and hence acts
as a catalyst for the whole cycle. The cycle commences once the stellar core temperature reaches 14 ×
106 K and is the primary source of energy in stars of mass M > 1.5 M⊙. Stars of lower mass convert
hydrogen to helium via an alternative process known as the ‘proton-proton chain’.

1. chemical that accelerates chemical reaction: a substance that increases the


rate of a chemical reaction without itself undergoing any change
2. stimulus to change: somebody or something that makes a change happen or
brings about an event
 The quarrel acted as a catalyst for the breakup of their partnership
TRIPLE ALPHA PROCESS
The combination or fusion of three alpha particles (helium nuclei 4He) to form a carbon nucleus (12C) is
known as the triple alpha process.

 The fusion process is not at all simple since there does not exist a stable configuration with an
atomic mass of 8 (8Be) that is formed by the fusion of two 4He nuclei. The lifetime of 8Be is a very short 3
× 10-16 seconds.
 However if this unstable 8Be nucleus is struck by another 4He nuclei it is possible to form 12C (and a gamma
ray). This can occur because the lifetime of 8Be is actually longer than the mean collision or scattering time
of helium nuclei at temperatures around 10 8K which are needed to make this fusion process proceed. In
1952 Edwin Salpeter proposed this process as the way to form 12C.
 Not long afterwards Fred Hoyle suggested that the fusion of 8Be and 4He would be greatly enhanced if
the 12C nucleus possessed an energy level near to the combined energies of the two nuclei involved. The
subsequent reaction would then be a resonant reaction. This resonant energy level was found by
experiments at the Kellogg Radiation Laboratory at CalTech.
The triple alpha process will occur in red giant stars that have left the main sequence (and have
consumed their core hydrogen) and have core temperatures of 108K and higher. Once 12C has been
formed it is possible with temperatures around 6 × 108 K to continue forming heavier nuclei by the
combination of two 12C nuclei to make 16O , 20Ne, 24Mg and with temperatures around 109K the
combination of two 16O nuclei can make 28Si, 31P, 31S and 32S

Neutron capture

Chart of nuclides showing thermal neutron capture cross section values

Science with neutrons

Neutron capture is a nuclear reaction in which an atomic nucleus and one or more neutrons collide and
merge to form a heavier nucleus.[1] Since neutrons have no electric charge, they can enter a nucleus
more easily than positively charged protons, which are repelled electrostatically.[1]
Neutron capture plays an important role in the cosmic nucleosynthesis of heavy elements. In stars it
can proceed in two ways: as a rapid (r-process) or a slow process (s-process).[1] Nuclei of masses greater
than 56 cannot be formed by thermonuclear reactions (i.e. by nuclear fusion), but can be formed by
neutron capture.[1] Neutron capture on protons yields a line at 2.223 MeV predicted[2] and commonly
observed[3] in solar flares.
Nuclear physics

Nuclear reaction
From Wikipedia, the free encyclopedia
Jump to navigationJump to search

Nucleus · Nucleons (p, n) · Nuclear


matter · Nuclear force · Nuclear
structure · Nuclear reaction

In this symbolic representing of a nuclear reaction, lithium-6 (6


3Li) and deuterium (2
1H) react to form the highly excited intermediate nucleus 8
4Be which then decays immediately into two alpha particles of helium-4 (4
2He
). Protons are symbolically represented by red spheres, and neutrons by blue spheres.
In nuclear physics and nuclear chemistry, a nuclear reaction is semantically considered to be the
process in which two nuclei, or else a nucleus of an atom and a subatomic particle (such as
a proton, neutron, or

high energy electron) from outside the atom, collide to produce one or more nuclides that are different
from the nuclide(s) that began the process (parent nuclei). Thus, a nuclear reaction must cause a
transformation of at least one nuclide to another. If a nucleus interacts with another nucleus or particle
and they then separate without changing the nature of any nuclide, the process is simply referred to as
a type of nuclear scattering, rather than a nuclear reaction.
In principle, a reaction can involve more than two particles colliding, but because the probability of
three or more nuclei to meet at the same time at the same place is much less than for two nuclei, such
an event is exceptionally rare (see triple alpha process for an example very close to a three-body
nuclear reaction). "Nuclear reaction" is a term implying an induced changing in a nuclide, and thus it
does not apply to any type of radioactive decay (which by definition is a spontaneous process).[citation
needed]

Natural nuclear reactions occur in the interaction between cosmic rays and matter, and nuclear
reactions can be employed artificially to obtain nuclear energy, at an adjustable rate, on demand.
Perhaps the most notable nuclear reactions are the nuclear chain reactions in fissionable materials that
produce induced nuclear fission, and the various nuclear fusion reactions of light elements that power
the energy production of the Sun and stars.
Atomic nucleus
Nuclear physics

Nucleus · Nucleons (p, n) · Nuclear

matter · Nuclea r
force · Nuclear structure · Nuclear reaction

A model of the atomic nucleus showing it as a compact bundle of the two types of nucleons: protons
(red) and neutrons (blue). In this diagram, protons and neutrons look like little balls stuck together, but
an actual nucleus (as understood by modern nuclear physics) cannot be explained like this, but only by
using quantum mechanics. In a nucleus which occupies a certain energy level (for example, the ground
state), each nucleon can be said to occupy a range of locations.
The atomic nucleus is the small, dense region consisting of protons and neutrons at the center of
an atom, discovered in 1911 by Ernest Rutherford based on the 1909 Geiger–Marsden gold foil
experiment. After the discovery of the neutron in 1932, models for a nucleus composed of protons and
neutrons were quickly developed by Dmitri Ivanenko[1] and Werner Heisenberg.[2][3][4][5][6] An atom is
composed of a positively-charged nucleus, with a cloud of negatively-charged electrons surrounding it,
bound together by electrostatic force. Almost all of the mass of an atom is located in the nucleus, with
a very small contribution from the electron cloud. Protons and neutrons are bound together to form a
nucleus by the nuclear force.
The diameter of the nucleus is in the range of 1.7566 fm (1.7566×10−15 m) for hydrogen (the diameter
of a single proton) to about 11.7142 fm for the heaviest atom uranium.[7] These dimensions are much
smaller than the diameter of the atom itself (nucleus + electron cloud), by a factor of about 26,634
(uranium atomic radius is about 156 pm (156×10−12 m))[8] to about 60,250 (hydrogen atomic radius is
about 52.92 pm).[a]
The branch of physics concerned with the study and understanding of the atomic nucleus, including its
composition and the forces which bind it together, is called nuclear physics.

Neutron

Neutron
The quark content of the neutron. The color
assignment of individual quarks is arbitrary, but
all three colors must be present. Forces between
quarks are mediated by gluons.

The neutron is a subatomic particle, symbol


n or n0
, with no net electric charge and a mass slightly greater than that of a proton. Protons and neutrons
constitute the nuclei of atoms. Since protons and neutrons behave similarly within the nucleus, and
each has a mass of approximately one atomic mass unit, they are both referred to as nucleons.[5] Their
properties and interactions are described by nuclear physics.
The chemical and nuclear properties of the nucleus are determined by the number of protons, called
the atomic number, and the number of neutrons, called the neutron number. The atomic mass
number is the total number of nucleons. For example, carbon has atomic number 6, and its
abundant carbon-12 isotope has 6 neutrons, whereas its rare carbon-13 isotope has 7 neutrons. Some
elements occur in nature with only one stable isotope, such as fluorine. Other elements occur with
many stable isotopes, such as tin with ten stable isotopes.
Within the nucleus, protons and neutrons are bound together through the nuclear force. Neutrons are
required for the stability of nuclei, with the exception of the single-proton hydrogen atom. Neutrons
are produced copiously in nuclear fission and fusion. They are a primary contributor to
the nucleosynthesis of chemical elements within stars through fission, fusion, and neutron
capture processes.
The neutron is essential to the production of nuclear power. In the decade after the neutron was
discovered by James Chadwick in 1932,[6] neutrons were used to induce many different types of nuclear
transmutations. With the discovery of nuclear fission in 1938,[7] it was quickly realized that, if a fission
event produced neutrons, each of these neutrons might cause further fission events, in a cascade
known as a nuclear chain reaction.[8] These events and findings led to the first self-sustaining nuclear
reactor (Chicago Pile-1, 1942) and the first nuclear weapon (Trinity, 1945).
Free neutrons, while not directly ionizing atoms, cause ionizing radiation. As such they can be a
biological hazard, depending upon dose.[8] A small natural "neutron background" flux of free neutrons
exists on Earth, caused by cosmic ray showers, and by the natural radioactivity of spontaneously
fissionable elements in the Earth's crust.[9] Dedicated neutron sources like neutron
generators, research reactors and spallation sources produce free neutrons for use in irradiation and
in neutron scattering experiments.

Proton

Proton
The quark content of a proton. The color
assignment of individual quarks is arbitrary, but
all three colors must be present. Forces between
quarks are mediated by gluons.

A proton is a subatomic particle, symbol


+
p or p
, with a positive electric charge of +1e elementary charge and a mass slightly less than that of a neutron.
Protons and neutrons, each with masses of approximately one atomic mass unit, are collectively
referred to as "nucleons".
One or more protons are present in the nucleus of every atom; they are a necessary part of the nucleus.
The number of protons in the nucleus is the defining property of an element, and is referred to as
the atomic number (represented by the symbol Z). Since each element has a unique number of
protons, each element has its own unique atomic number.
The word proton is Greek for "first", and this name was given to the hydrogen nucleus by Ernest
Rutherford in 1920. In previous years, Rutherford had discovered that the hydrogen nucleus (known to
be the lightest nucleus) could be extracted from the nuclei of nitrogen by atomic collisions.[4] Protons
were therefore a candidate to be a fundamental particle, and hence a building block of nitrogen and all
other heavier atomic nuclei.
Although protons were originally considered fundamental or elementary particles, in the
modern Standard Model of particle physics, protons are classified as hadrons, like neutrons, the
other nucleon (particles present in atomic nuclei), composite particles composed of three valence
quarks: two up quarks of charge +2/3e and one down quark of charge –1/3e. The rest masses of quarks
contribute only about 1% of a proton's mass, however.[5] The remainder of a proton's mass is due
to quantum chromodynamics binding energy, which includes the kinetic energy of the quarks and the
energy of the gluon fields that bind the quarks together. Because protons are not fundamental
particles, they possess a measurable size; the root mean square charge radius of a proton is about 0.84–
0.87 fm or 0.84×10−15 to 0.87×10−15 m.[6][7]
At sufficiently low temperatures, free protons will bind to electrons. However, the character of such
bound protons does not change, and they remain protons. A fast proton moving through matter will
slow by interactions with electrons and nuclei, until it is captured by the electron cloud of an atom. The
result is a protonated atom, which is a chemical compound of hydrogen. In vacuum, when free
electrons are present, a sufficiently slow proton may pick up a single free electron, becoming a
neutral hydrogen atom, which is chemically a free radical. Such "free hydrogen atoms" tend to react
chemically with many other types of atoms at sufficiently low energies. When free hydrogen atoms
react with each other, they form neutral hydrogen molecules (H 2), which are the most common
molecular component of molecular clouds in interstellar space.

Electrostatics
Part of a series of articles about

Electromagnetism
Electrostatics is a branch of physics that studies electric charges at rest.
An electrostatic effect: styrofoam peanuts clinging to a cat's fur due to static electricity.
The triboelectric effect causes an electrostatic charge to build up on the surface of the fur due to the
cat's motions. The electric field of the charge causes polarization of the molecules of the styrofoam due
to electrostatic induction, resulting in a slight attraction of the light plastic pieces to the charged fur.
This effect is also the cause of static cling in clothes.
Since classical physics, it has been known that some materials such as amber attract lightweight
particles after rubbing. The Greek word for amber, ήλεκτρον, or electron, was the source of the word
'electricity'. Electrostatic phenomena arise from the forces that electric charges exert on each other.
Such forces are described by Coulomb's law. Even though electrostatically induced forces seem to be
rather weak, some electrostatic forces such as the one between an electron and a proton, that together
make up a hydrogen atom, is about 36 orders of magnitude stronger than the gravitational force acting
between them.
There are many examples of electrostatic phenomena, from those as simple as the attraction of the
plastic wrap to one's hand after it is removed from a package to the apparently spontaneous explosion
of grain silos, the damage of electronic components during manufacturing, and photocopier & laser
printer operation. Electrostatics involves the buildup of charge on the surface of objects due to contact
with other surfaces. Although charge exchange happens whenever any two surfaces contact and
separate, the effects of charge exchange are usually only noticed when at least one of the surfaces has
a high resistance to electrical flow. This is because the charges that transfer are trapped there for a
time long enough for their effects to be observed. These charges then remain on the object until they
either bleed off to ground or are quickly neutralized by a discharge: e.g., the familiar phenomenon of a
static 'shock' is caused by the neutralization of charge built up in the body from contact with insulated
surfaces.

Nucleosynthesis
"Nucleogenesis" redirects here. For the song by Vangelis, see Albedo 0.39.
Nucleosynthesis is the process that creates new atomic nuclei from pre-existing nucleons (protons and
neutrons). The first nuclei were formed a few minutes after the Big Bang, through the process called Big
Bang nucleosynthesis. After about 20 minutes, the universe had cooled to a point at which these
processes ended, so only the fastest and simplest reactions occurred, leaving our universe containing
about 75% hydrogen, 24% helium by mass. The rest is traces of other elements such as lithium and the
hydrogen isotope deuterium. The universe still has approximately the same composition.
Stars fuse light elements to heavier ones in their cores, giving off energy in the process known as stellar
nucleosynthesis. Fusion processes create many of the lighter elements, up to and
including iron and nickel in the most massive stars although these mostly remain trapped in stellar
cores and remnants. The s-process creates heavy elements, from strontium upwards.
Supernova nucleosynthesis within exploding stars is largely responsible for the elements
between oxygen and rubidium: from the ejection of elements produced during stellar nucleosynthesis;
through explosive nucleosynthesis during the supernova explosion; and from the r-process (absorption
of multiple neutrons) during the explosion.
Neutron star mergers are also very responsible for the synthesis of many heavy elements, via the r-
process ("r" stands for "rapid"). When two neutron stars collide, a large amount of neutron-rich matter
may be ejected at extremely high temperatures, and very heavy elements form as the ejecta begins to
cool.
Cosmic ray spallation, caused when cosmic rays impact the interstellar medium and fragment larger
atomic species, is a significant source of the lighter nuclei, particularly 3He, 9Be and 10,11B, that are not
created by stellar nucleosynthesis. Cosmic ray bombardment of elements on Earth also contribute to
the presence of rare, short-lived atomic species called cosmogenic nuclides.
In addition to the fusion processes responsible for the growing abundances of elements in the universe,
a few minor natural processes continue to produce very small numbers of new nuclides on Earth. These
nuclides contribute little to their abundances, but may account for the presence of specific new nuclei.
These nuclides are produced via radiogenesis (decay) of long-lived, heavy, primordial radionuclides
such as uranium and thorium.

r-process
Nuclear physics

Nucleus · Nucleons (p, n) · Nuclear


matter · Nuclear force · Nuclear
structure · Nuclear reaction

Schematic illustrating the r-process as it occurs in supernovae or neutron star collisions.[1] Neutrons are
rapidly absorbed faster than the resulting nuclei can beta-decay; this allows the r-process to produce
very neutron-rich nuclei following the neutron drip line. There are waiting points located at magic
numbers N = 50, 82, 126, where beta-decay is favored due to low neutron-capture cross sections
resulting from the closed shells. The cycle then repeats until the next waiting point, creating yet heavier
nuclei of elements up to the actinides; the natural abundance of these elements results entirely from
the r-process. In the superheavy mass region (A ≥ 270), neutron-induced fission or spontaneous
fission are expected to become dominant and end the r-process. Most nuclei along the hypothetical r-
process path, however, are unknown.[2]
The rapid neutron-capture process, or so-called r-process, is a set of nuclear reactions that in nuclear
astrophysics is responsible for the creation of approximately half of the atomic nuclei heavier than iron;
the "heavy elements", with the other half produced by the p-process and s-process. The r-process
usually synthesizes the most neutron-rich stable isotopes of each heavy element. The r-process can
typically synthesize the heaviest four isotopes of every heavy element, and the two heaviest isotopes,
which are referred to as r-only nuclei, can only be created via the r-process. Abundance peaks for the r-
process occur near atomic weights A = 82 (elements Se, Br and Kr), A = 130 (elements Te, I, and Xe)
and A = 196 (elements Os, Ir and Pt).
The r-process entails a succession of rapid neutron captures (hence the name) by one or more
heavy seed nuclei, typically beginning with nuclei in the abundance peak centered on 56Fe. The captures
must be rapid in the sense that the nuclei must not have time to undergo radioactive decay (typically
via β- decay) before another neutron arrives to be captured. This sequence can continue up to the limit
of stability of the increasingly neutron-rich nuclei (the neutron drip line) to physically retain neutrons
as governed by the short range nuclear force. The r-process therefore must occur in locations where
there exist a high density of free neutrons. Early studies theorized that 1024 free neutrons per
cm3 would be required, for temperatures about 1GK, in order to match the waiting points, at which no
more neutrons can be captured, with the atomic numbers of the abundance peaks for r-process
nuclei.[3] This amounts to almost a gram of free neutrons in every cubic centimeter, an astonishing
number requiring extreme locations.[a] Traditionally this suggested the material ejected from the
reexpanded core of a core-collapse supernova, as part of supernova nucleosynthesis,[4] or
decompression of neutron-star matter thrown off by a binary neutron star merger.[5] The relative
contributions of these sources to the astrophysical abundance of r-process elements is a matter of
ongoing research.[6]
A limited r-process-like series of neutron captures occurs to a minor extent in thermonuclear
weapon explosions. These led to the discovery of the elements einsteinium (element 99)
and fermium (element 100) in nuclear weapon fallout.
The r-process contrasts with 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 ordinary stars, particularly AGB stars, where the neutron flux is sufficient to
cause neutron captures to recur every 10–100 years, much too slow for the r-process, which requires
100 captures per second. The s-process is secondary, meaning that it requires pre-existing heavy
isotopes as seed nuclei to be converted into other heavy nuclei by a slow sequence of captures of free
neutrons. The r-process scenarios create their own seed nuclei, so they might proceed in massive stars
that contain no heavy seed nuclei. Taken together, the r- and s-processes account for almost the
entire abundance of chemical elements heavier than iron. The historical challenge has been to locate
physical settings appropriate for their time scales.

s-process
Jump to navigationJump to search
The slow neutron-capture process , or s-process is a series of reactions in nuclear astrophysics that
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 table of nuclides.
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 with
higher fluxes of free neutrons; it produces heavier elements and more neutron-rich isotopes than the s-
process. Together the two processes account for most of the relative abundance of chemical
elements heavier than iron.

Mass number
Nuclear physics
Nucleus · Nucleons (p, n) · Nuclear
matter · Nuclear force · Nuclear
structure · Nuclear reaction

The mass number (symbol A, from the German word Atomgewicht (atomic weight)),[1] also
called atomic mass number or nucleon number, is the total number of protons and neutrons (together
known as nucleons) in an atomic nucleus. It is approximately equal to the atomic (also known
as isotopic) mass of the atom expressed in atomic mass units. Because protons and neutrons both
are baryons, the mass number A is identical with the baryon number B as of the nucleus as of the whole
atom or ion. The mass number is different for each different isotope of a chemical element. Hence, the
difference between the mass number and the atomic number Z gives the number of neutrons (N) in a

given nucleus: .[2]


The mass number is written either after the element name or as a superscript to the left of an element's
symbol. For example, the most common isotope of carbon is carbon-12, or 12
C
, which has 6 protons and 6 neutrons. The full isotope symbol would also have the atomic number (Z)
as a subscript to the left of the element symbol directly below the mass number: 12
6C
.[3] This is technically redundant, as each element is defined by its atomic number, so it is often omitted.

Iron peak
From Wikipedia, the free encyclopedia
Jump to navigationJump to search
The iron peak is a local maximum in the vicinity of Fe (Cr, Mn, Fe, Co and Ni) on the graph of
the abundances of the chemical elements.
For elements lighter than iron on the periodic table, nuclear fusion releases energy. For iron, and for all
of the heavier elements, nuclear fusion consumes energy. Chemical elements up to the iron peak are
produced in ordinary stellar nucleosynthesis, with the alpha elements being particular abundant. Some
heavier elements are produced by less efficient processes such as the r-process and s-process.
Elements with atomic numbers close to iron are produced in large quantities in supernova due to
explosive oxygen and silicon fusion, followed by radioactive decay of nuclei such as Nickel-56. On
average, heavier elements are less abundant in the universe, but some of those near iron are
comparatively more abundant than would be expected from this trend.[1]
Abundances of the chemical elements in the Solar System. Hydrogen and helium are most common,
from the Big Bang. The next three elements (Li, Be, B) are rare because they are poorly synthesized in
the Big Bang and also in stars. The two general trends in the remaining stellar-produced elements are:
(1) an alternation of abundance in elements as they have even or odd atomic numbers, and (2) a general
decrease in abundance, as elements become heavier. The "iron peak" may be seen in the elements
near iron as a secondary effect, increasing relative abundances of elements with nuclei most strongly
bound.
A graph of the nuclear binding energy per nucleon for all the elements shows a sharp increase to a peak
near nickel and then a slow decrease to heavier elements. Increasing values of binding energy represent
energy released when a collection of nuclei is rearranged into another collection for which the sum of
nuclear binding energies is higher. Light elements such as hydrogen release large amounts of energy (a
big increase in binding energy) when combined to form heavier nuclei. Conversely, heavy elements
such as uranium release energy when converted to lighter nuclei though alpha decay and nuclear
fission. 56
28Ni
is the most thermodynamically favorable in the cores of high-mass stars. Although iron-58 and nickel-
62 have even higher (per nucleon) binding energy, their synthesis cannot be achieved in large
quantities because the required number of neutrons is typically not available in the stellar nuclear
material.
Nuclear fusion
From Wikipedia, the free encyclopedia
This article is about the science behind nuclear fusion. For its use in producing energy, see Fusion power.
The Sun is a main-sequence star, and thus generates its energy by nuclear fusion of hydrogen nuclei
into helium. In its core, the Sun fuses 60 million metric tons of hydrogen each second.

The nuclear binding energy curve. The formation of nuclei with masses up to Iron-56 releases energy,
as illustrated above.

Nuclear physics
Nucleus · Nucleons (p, n) · Nuclear
matter · Nuclear force · Nuclear
structure · Nuclear reaction

In nuclear physics and nuclear chemistry, nuclear fusion is a reaction in which two or more atomic
nuclei are combined to form one or more different atomic nuclei and subatomic particles
(neutrons or protons). The difference in mass between the reactants and products is manifested as
either the release or absorption of energy. This difference in mass arises due to the difference in atomic
"binding energy" between the atomic nuclei before and after the reaction. Fusion is the process that
powers active or "main sequence" stars, or other high magnitude stars.
A fusion process that produces a nucleus lighter than iron-56 or nickel-62 will generally yield a net
energy release. These elements have the smallest mass per nucleon and the largest binding
energy per nucleon, respectively. Fusion of light elements toward these releases energy
(an exothermic process), while a fusion producing nuclei heavier than these elements will result in
energy retained by the resulting nucleons, and the resulting reaction is endothermic. The opposite is
true for the reverse process, nuclear fission. This means that the lighter elements, such as hydrogen
and helium, are in general more fusible; while the heavier elements, such
as uranium, thorium and plutonium, are more fissionable. The extreme astrophysical event of
a supernova can produce enough energy to fuse nuclei into elements heavier than iron.
In 1920, Arthur Eddington suggested hydrogen-helium fusion could be the primary source of stellar
energy. Quantum tunneling was discovered by Friedrich Hund in 1929, and shortly afterwards Robert
Atkinson and Fritz Houtermans used the measured masses of light elements to show that large
amounts of energy could be released by fusing small nuclei. Building on the early experiments in nuclear
transmutation by Ernest Rutherford, laboratory fusion of hydrogen isotopes was accomplished
by Mark Oliphant in 1932. In the remainder of that decade, the theory of the main cycle of nuclear
fusion in stars were worked out by Hans Bethe. Research into fusion for military purposes began in the
early 1940s as part of the Manhattan Project. Fusion was accomplished in 1951 with the Greenhouse
Item nuclear test. Nuclear fusion on a large scale in an explosion was first carried out on 1 November
1952, in the Ivy Mike hydrogen bomb test.
Research into developing controlled fusion inside fusion reactors has been ongoing since the 1940s, but
the technology is still in its development phase.

Solar flare
"Sun flare" redirects here. For the rose variety, see Rosa 'Sun Flare'.

Solar flare and its prominence eruption recorded on June 7, 2011 by SDO in extreme ultraviolet

Evolution of magnetism on the Sun.

On August 31, 2012 a long prominence/filament of solar material that had been hovering in the Sun's
atmosphere, the corona, erupted out into space at 4:36 p.m. EDT. Seen here from the Solar Dynamics
Observatory, the flare caused on Earth on September 3.
A solar flare is a sudden flash of increased brightness on the Sun, usually observed near its surface and
in close proximity to a sunspot group. Powerful flares are often, but not always, accompanied by
a coronal mass ejection. Even the most powerful flares are barely detectable in the total solar
irradiance (the "solar constant").[1]
Solar flares occur in a power-law spectrum of magnitudes; an energy release of typically
1020 joules of energy suffices to produce a clearly observable event, while a major event can emit up to
1025 joules.[2]
Flares are closely associated with the ejection of plasmas and particles through
the Sun's corona into outer space; flares also copiously emit radio waves. If the ejection is in the
direction of the Earth, particles associated with this disturbance can penetrate into the upper
atmosphere (the ionosphere) and cause bright auroras, and may even disrupt long range radio
communication. It usually takes days for the solar plasma ejecta to reach Earth.[3] Flares also occur on
other stars, where the term stellar flare applies. High-energy particles, which may be relativistic, can
arrive almost simultaneously with the electromagnetic radiations.
On July 23, 2012, a massive, potentially damaging, solar storm (solar flare, coronal mass ejection
and electromagnetic radiation) barely missed Earth.[4][5] In 2014, Pete Riley of Predictive Science Inc.
published a paper in which he attempted to calculate the odds of a similar solar storm hitting Earth
within the next 10 years, by extrapolating records of past solar storms from the 1960s to the present
day. He concluded that there may be as much as a 12% chance of such an event occurring.[4]

Вам также может понравиться