Neutron

Source: Wikipedia, the free encyclopedia.
Neutron
Mean lifetime
879.4(6) s (free)[5]
Electric chargee
(−2±8)×10−22 
) = 1/2(1/2+)

The neutron is a subatomic particle, symbol
n
or
n0
, which has a neutral (not positive or negative) 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, they are both referred to as nucleons. Nucleons have a mass of approximately one atomic mass unit, or dalton, symbol Da. Their properties and interactions are described by nuclear physics. Protons and neutrons are not elementary particles; each is composed of three quarks.

The

chemical properties of an atom are mostly determined by the configuration of electrons that orbit the atom's heavy nucleus. The electron configuration is determined by the charge of the nucleus, which is determined by the number of protons, or atomic number. The number of neutrons is the neutron number
. Neutrons do not affect the electron configuration.

Atoms of a

isotopes. For example, carbon, with atomic number 6, has an abundant isotope carbon-12 with 6 neutrons and a rare isotope carbon-13 with 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, or with no stable isotope, such as technetium
.

The properties of an atomic nucleus depend on both atomic and neutron numbers. With their positive charge, the protons within the nucleus are repelled by the long-range

electromagnetic force, but the much stronger, but short-range, nuclear force binds the nucleons closely together. Neutrons are required for the stability of nuclei, with the exception of the single-proton hydrogen nucleus. 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, neutrons were used to induce many different types of nuclear transmutations. With the discovery of nuclear fission in 1938, 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. These events and findings led to the first self-sustaining nuclear reactor (Chicago Pile-1, 1942) and the first nuclear weapon (Trinity, 1945).

Dedicated

antineutrino, with a mean lifetime of about 15 minutes. Free neutrons do not directly ionize atoms, but they do indirectly cause ionizing radiation, so they can be a biological hazard, depending on dose. 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
.

Neutrons in an atomic nucleus

An

atomic mass number, A, is equal to the sum of atomic and neutron numbers. Nuclides with the same atomic mass number, but different atomic and neutron numbers, are called isobars. The mass of a nucleus is always slightly less than the sum of its proton and neutron masses: the difference in mass represents the mass equivalent to nuclear binding energy, the energy which would need to be added to take the nucleus apart.[8]
: 822 

The nucleus of the most common isotope of the hydrogen atom (with the chemical symbol 1H) is a lone proton. The nuclei of the heavy hydrogen isotopes deuterium (D or 2H) and tritium (T or 3H) contain one proton bound to one and two neutrons, respectively. All other types of atomic nuclei are composed of two or more protons and various numbers of neutrons. The most common nuclide of the common chemical element lead, 208Pb, has 82 protons and 126 neutrons, for example. The table of nuclides comprises all the known nuclides. Even though it is not a chemical element, the neutron is included in this table.[9]

Nuclear fission caused by absorption of a neutron by uranium-235. The heavy nuclide fragments into lighter components and additional neutrons.

Protons and neutrons behave almost identically under the influence of the nuclear force within the nucleus. They are therefore both referred to collectively as nucleons.[10] The concept of isospin, in which the proton and neutron are viewed as two quantum states of the same particle, is used to model the interactions of nucleons by the nuclear or weak forces. Because of the strength of the nuclear force at short distances, the nuclear energy binding nucleons is more than seven orders of magnitude larger than the electromagnetic energy binding electrons in atoms. Nuclear reactions (such as nuclear fission) therefore have an energy density that is more than ten million times that of chemical reactions. Ultimately, the ability of the nuclear force to store energy arising from the electromagnetic repulsion of nuclear components is the basis for most of the energy that makes nuclear reactors or bombs possible. In nuclear fission, the absorption of a neutron by a heavy nuclide (e.g., uranium-235) causes the nuclide to become unstable and break into light nuclides and additional neutrons. The positively charged light nuclides then repel, releasing electromagnetic potential energy.

Beta decay

Neutrons and protons within a nucleus behave similarly and can exchange their identities by similar reactions. These reactions are a form of radioactive decay known as beta decay. Beta decay, in which neutrons decay to protons, or vice versa, is governed by the weak force, and it requires the emission or absorption of electrons and neutrinos, or their antiparticles. The neutron and proton decay reactions are:


n0

p+
+
e
+
ν
e

where
p+
,
e
, and
ν
e
denote the proton, electron and electron anti-neutrino decay products,[11] and


p+

n0
+
e+
+
ν
e

where
n0
,
e+
, and
ν
e
denote the neutron, positron and electron neutrino decay products.

The electron and positron produced in these reactions are historically known as

beta particles
, denoted β or β+ respectively, lending the name to the decay process. In these reactions, the original particle is not composed of the product particles; rather, the product particles are created at the instant of the reaction.

The "free" neutron

"Free" neutrons or protons are nucleons that exist independently, free of any nucleus.

The free neutron has a mass of 939565413.3 

magnetic moment, so the neutron is influenced by magnetic fields. The specific properties of the neutron are described below in the Intrinsic properties section
.

Outside the nucleus, free neutrons are unstable and have a

mean lifetime of 879.6±0.8 s (about 14 minutes, 40 seconds) by beta decay; therefore the half-life for this process (which differs from the mean lifetime by a factor of ln(2) = 0.693) is 610.1±0.7 s (about 10 minutes, 10 seconds).[14][15]
This decay, which produces a proton, an electron and electron anti-neutrino, is possible because the mass of the neutron is slightly greater than that of the proton. By the mass-energy equivalence, when a neutron decays to a proton this way, a lower energy state is attained.

For the free neutron the decay energy for this process (based on the masses of the neutron, proton, and electron) is 0.782343 MeV. By comparison the mass energy of the neutron is 939.6 MeV. The maximal energy of the beta decay electron (in the process wherein the neutrino receives a vanishingly small amount of kinetic energy) has been measured at 0.782±0.013 MeV.[16] The latter number is not well-enough measured to determine the comparatively tiny rest mass of the neutrino (which must in theory be subtracted from the maximal electron kinetic energy). The neutrino mass is better constrained by many other methods.

The decay of a free proton to a more massive neutron is energetically disallowed. A high-energy collision of a proton and an electron or neutrino can result in a neutron, however.

A small fraction (about one in 1000) of free neutrons decay with the same products, but add an extra particle in the form of an emitted gamma ray:


n0

p+
+
e
+
ν
e
+
γ

This gamma ray may be thought of as an "internal bremsstrahlung" that arises from the electromagnetic interaction of the emitted beta particle with the proton. Internal bremsstrahlung gamma ray production is also a minor feature of beta decays of bound neutrons (as discussed below).

A schematic of the nucleus of an atom indicating
β
radiation, the emission of a fast electron from the nucleus (the accompanying antineutrino is omitted). In the Rutherford model for the nucleus, red spheres were protons with positive charge and blue spheres were protons tightly bound to an electron with no net charge.
The inset shows beta decay of a free neutron as it is understood today; an electron and antineutrino are created in this process.

A very small minority of neutron decays (about four per million) are so-called "two-body (neutron) decays", in which a proton, electron and antineutrino are produced as usual, but the electron fails to gain the 13.6 eV necessary energy to escape the proton (the ionization energy of hydrogen), and therefore simply remains bound to it, forming a neutral hydrogen atom (one of the "two bodies"). In this type of free neutron decay, almost all of the neutron decay energy is carried off by the antineutrino (the other "body"). (The hydrogen atom recoils with a speed of only about (decay energy)/(hydrogen rest energy) times the speed of light, or 250 km/s.)

Neutrons and protons bound in a nucleus

Neutrons are a necessary constituent of any atomic nucleus that contains more than one proton. As a result of their positive charges, interacting protons have a mutual

diproton and neutron–proton ratio).[17] Neutrons bind with protons and one another in the nucleus via the nuclear force
, effectively moderating the repulsive forces between the protons and stabilizing the nucleus.

While a free neutron has a half life of about 10.2 min and a free proton is stable, within nuclei neutrons are often stable and protons are sometimes unstable. When bound within a nucleus, nucleons can

quantum numbers
. Nucleon decay within a nucleus can occur if allowed by basic energy conservation and quantum mechanical constraints. The decay products, that is, the emitted particles, carry away the energy excess as a nucleon falls from one quantum state to one with less energy, while the neutron (or proton) changes to a proton (or neutron).

For a neutron to decay, the resulting proton requires an available state at lower energy than the initial neutron state. In stable nuclei the possible lower energy states are all filled, meaning each state is occupied by a pair of protons, one with spin up, another with spin down. When all available proton states are filled, the Pauli exclusion principle disallows the decay of a neutron to a proton within stable nuclei. The situation is similar to electrons of an atom, where electrons that occupy distinct atomic orbitals are prevented by the exclusion principle from decaying to lower, already-filled, energy states, with the emission of a photon. The stability of nuclei and nuclide radioactivity are consequences of these constraints.

One example of the decay of a neutron within a nuclide is the

nitrogen-14
(7 protons, 7 neutrons) with a half-life of about 5,730 years. The decay emits an electron and an electron anti-neutrino. Nitrogen-14 is stable, since none of its protons or neutrons have available quantum states of lesser energy.

The transformation of a proton to a neutron inside of a nucleus is also possible through electron capture:


p+
+
e

n0
+
ν
e

Positron capture by neutrons in nuclei that contain an excess of neutrons is also possible, but is hindered because positrons are both relatively rare in ordinary matter and quickly annihilate when they encounter electrons (which are much less rare) and in any case are repelled by the positive nucleus. Similar, but far more rare, reactions involve the capture of a neutrino by a nucleon in inverse beta decay.

Competition of beta decay types

Three types of beta decay in competition are illustrated by the single isotope copper-64 (29 protons, 35 neutrons), which has a half-life of about 12.7 hours. This isotope has one unpaired proton and one unpaired neutron, so either the proton or the neutron can decay. This particular nuclide is almost equally likely to undergo proton decay (by positron emission, 18% or by electron capture, 43%; both forming 64
Ni
) or neutron decay (by electron emission, 39%; forming 64
Zn
).

The neutron in elementary particle physics - the Standard Model


W
boson
.
The principal Feynman diagram for
β+
 decay of a proton into a neutron, positron, and electron neutrino via an intermediate heavy
W+
boson.

Within the theoretical framework of Standard Model for particle physics, a neutron comprises two

composite, rather than elementary
, particle.

The quarks of the neutron are held together by the strong force, mediated by gluons.[19] The nuclear force results from secondary effects of the more fundamental strong force.

The only possible decay mode for the neutron that

W boson. By this process, the Standard Model description of beta decay, the neutron decays into a proton (which contains one down and two up quarks), an electron, and an electron antineutrino
.

The decay of the proton to a neutron occurs similarly through the weak force. The decay of one of the proton's up quarks into a down quark can be achieved by the emission of a W boson. The proton decays into a neutron, a positron, and an electron neutrino. This reaction can only occur within an atomic nucleus which has a quantum state at lower energy available for the created neutron.

Discovery

The story of the discovery of the neutron and its properties is central to the extraordinary developments in atomic physics that occurred in the first half of the 20th century, leading ultimately to the atomic bomb in 1945. In the 1911 Rutherford model, the atom consisted of a small positively charged massive nucleus surrounded by a much larger cloud of negatively charged electrons. In 1920, Ernest Rutherford suggested that the nucleus consisted of positive protons and neutrally charged particles, suggested to be a proton and an electron bound in some way.[20] Electrons were assumed to reside within the nucleus because it was known that beta radiation consisted of electrons emitted from the nucleus.[20] About the time Rutherford suggested the neutral proton-electron composite, several other publications appeared making similar suggestions, and in 1921 the American chemist W. D. Harkins first named the hypothetical particle a "neutron".[21][22] The name derives from the Latin root for neutralis (neuter) and the Greek suffix -on (a suffix used in the names of subatomic particles, i.e. electron and proton).[23][24] References to the word neutron in connection with the atom can be found in the literature as early as 1899, however.[22]

Throughout the 1920s, physicists assumed that the atomic nucleus was composed of protons and "nuclear electrons",

Heisenberg uncertainty relation of quantum mechanics.[27][28] The Klein paradox,[29] discovered by Oskar Klein in 1928, presented further quantum mechanical objections to the notion of an electron confined within a nucleus.[27]
Observed properties of atoms and molecules were inconsistent with the nuclear spin expected from the proton–electron hypothesis. Both protons and electrons carry an intrinsic spin of 1/2ħ. Isotopes of the same species (i.e. having the same number of protons) can have both integer or fractional spin, i.e. the neutron spin must be also fractional (1/2ħ). But there is no way to arrange the spins of an electron and a proton (supposed to bond to form a neutron) to get the fractional spin of a neutron.

In 1931,

gamma radiation.[30][31] The following year Irène Joliot-Curie and Frédéric Joliot-Curie in Paris showed that if this "gamma" radiation fell on paraffin, or any other hydrogen-containing compound, it ejected protons of very high energy.[32] Neither Rutherford nor James Chadwick at the Cavendish Laboratory in Cambridge were convinced by the gamma ray interpretation.[33] Chadwick quickly performed a series of experiments that showed that the new radiation consisted of uncharged particles with about the same mass as the proton.[34][35][36] These properties matched Rutherford's hypothesized neutron. Chadwick won the 1935 Nobel Prize in Physics for this discovery.[2]

Models depicting the nucleus and electron energy levels in hydrogen, helium, lithium, and neon atoms. In reality, the diameter of the nucleus is about 100,000 times smaller than the diameter of the atom.

Models for an atomic nucleus consisting of protons and neutrons were quickly developed by Werner Heisenberg[37][38][39] and others.[40][41] The proton–neutron model explained the puzzle of nuclear spins. The origins of beta radiation were explained by Enrico Fermi in 1934 by the process of beta decay, in which the neutron decays to a proton by creating an electron and a (at the time undiscovered) neutrino.[42] In 1935, Chadwick and his doctoral student Maurice Goldhaber reported the first accurate measurement of the mass of the neutron.[43][44]

By 1934, Fermi had bombarded heavier elements with neutrons to induce radioactivity in elements of high atomic number. In 1938, Fermi received the Nobel Prize in Physics "for his demonstrations of the existence of new radioactive elements produced by neutron irradiation, and for his related discovery of nuclear reactions brought about by slow neutrons".[45] In December 1938 Otto Hahn, Lise Meitner, and Fritz Strassmann discovered nuclear fission, or the fractionation of uranium nuclei into lighter elements, induced by neutron bombardment.[46][47][48][49] In 1945 Hahn received the 1944 Nobel Prize in Chemistry "for his discovery of the fission of heavy atomic nuclei".[50][51][52]

The discovery of nuclear fission would lead to the development of nuclear power and the atomic bomb by the end of World War II. 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.[7] These events and findings led Fermi to construct the Chicago Pile-1 at the University of Chicago in 1942, the first self-sustaining nuclear reactor. Just three years later the Manhattan Project was able to test the first atomic bomb, the Trinity nuclear test in July 1945.

Properties

Mass

The mass of a neutron cannot be directly determined by

deuteron can be measured with a mass spectrometer, the mass of a neutron can be deduced by subtracting proton mass from deuteron mass, with the difference being the mass of the neutron plus the binding energy
of deuterium (expressed as a positive emitted energy). The latter can be directly measured by measuring the energy () of the single 2.224 MeV gamma photon emitted when a deuteron is formed by a proton capturing a neutron (this is exothermic and happens with zero-energy neutrons). The small recoil kinetic energy () of the deuteron (about 0.06% of the total energy) must also be accounted for.

The energy of the gamma ray can be measured to high precision by X-ray diffraction techniques, as was first done by Bell and Elliot in 1948. The best modern (1986) values for neutron mass by this technique are provided by Greene, et al.[53] These give a neutron mass of:

mneutron = 1.008644904(14) Da

The value for the neutron mass in MeV is less accurately known, due to less accuracy in the known conversion of Da to MeV/c2:[54]

mneutron = 939.56563(28) MeV/c2.

Another method to determine the mass of a neutron starts from the beta decay of the neutron, when the momenta of the resulting proton and electron are measured.

Spin

The neutron is a spin 1/2 particle, that is, it is a

Dirac particle, the possibility that the neutron was a spin 3/2 particle lingered. The interactions of the neutron's magnetic moment with an external magnetic field were exploited to finally determine the spin of the neutron.[55] In 1949, Hughes and Burgy measured neutrons reflected from a ferromagnetic mirror and found that the angular distribution of the reflections was consistent with spin 1/2.[56] In 1954, Sherwood, Stephenson, and Bernstein employed neutrons in a Stern–Gerlach experiment that used a magnetic field to separate the neutron spin states. They recorded two such spin states, consistent with a spin 1/2 particle.[55][57]

As a fermion, the neutron is subject to the

degeneracy pressure which counteracts gravity in neutron stars and prevents them from forming black holes.[58]

Magnetic moment

Even though the neutron is a neutral particle, the magnetic moment of a neutron is not zero. The neutron is not affected by electric fields, but it is affected by magnetic fields. The value for the neutron's magnetic moment was first directly measured by Luis Alvarez and Felix Bloch at Berkeley, California, in 1940.[59] Alvarez and Bloch determined the magnetic moment of the neutron to be μn= −1.93(2) μN, where μN is the nuclear magneton. The neutron's magnetic moment has a negative value, because its orientation is opposite to the neutron's spin.[60]

The magnetic moment of the neutron is an indication of its quark substructure and internal charge distribution.

hadrons, the neutron is composed of one up quark (charge +2/3 e) and two down quarks (charge −1/3 e).[61] The magnetic moment of the neutron can be modeled as a sum of the magnetic moments of the constituent quarks.[62]
The calculation assumes that the quarks behave like pointlike Dirac particles, each having their own magnetic moment. Simplistically, the magnetic moment of the neutron can be viewed as resulting from the vector sum of the three quark magnetic moments, plus the orbital magnetic moments caused by the movement of the three charged quarks within the neutron.

In one of the early successes of the Standard Model, in 1964 Mirza A.B. Beg, Benjamin W. Lee, and Abraham Pais calculated the ratio of proton to neutron magnetic moments to be −3/2 (or a ratio of −1.5), which agrees with the experimental value to within 3%.[63][64][65] The measured value for this ratio is −1.45989805(34).[4]

The above treatment compares neutrons with protons, allowing the complex behavior of quarks to be subtracted out between models, and merely exploring what the effects would be of differing quark charges (or quark type). Such calculations are enough to show that the interior of neutrons is very much like that of protons, save for the difference in quark composition with a down quark in the neutron replacing an up quark in the proton.

The neutron magnetic moment can be roughly computed by assuming a simple

wavefunction for baryons composed of three quarks. A straightforward calculation gives fairly accurate estimates for the magnetic moments of neutrons, protons, and other baryons.[62]
For a neutron, the result of this calculation is that the magnetic moment of the neutron is given by μn= 4/3 μd − 1/3 μu, where μd and μu are the magnetic moments for the down and up quarks, respectively. This result combines the intrinsic magnetic moments of the quarks with their orbital magnetic moments, and assumes the three quarks are in a particular, dominant quantum state.

Baryon Magnetic moment
of quark model
Computed
()
Observed
()
p 4/3 μu − 1/3 μd 2.79 2.793
n 4/3 μd − 1/3 μu −1.86 −1.913

The results of this calculation are encouraging, but the masses of the up or down quarks were assumed to be 1/3 the mass of a nucleon.

strong force.[66][67] Furthermore, the complex system of quarks and gluons that constitute a neutron requires a relativistic treatment.[68] But the nucleon magnetic moment has been successfully computed numerically from first principles, including all of the effects mentioned and using more realistic values for the quark masses. The calculation gave results that were in fair agreement with measurement, but it required significant computing resources.[69][70]

Electric charge

The total electric charge of the neutron is e. This zero value has been tested experimentally, and the present experimental limit for the charge of the neutron is −2(8)×10−22 e,[6] or −3(13)×10−41 C. This value is consistent with zero, given the experimental uncertainties (indicated in parentheses). By comparison, the charge of the proton is +1 e.

Structure and geometry of charge distribution

An article published in 2007 featuring a model-independent analysis concluded that the neutron has a negatively charged exterior, a positively charged middle, and a negative core.[71] In a simplified classical view, the negative "skin" of the neutron assists it to be attracted to the protons with which it interacts in the nucleus; but the main attraction between neutrons and protons is via the nuclear force, which does not involve electric charge.

The simplified classical view of the neutron's charge distribution also "explains" the fact that the neutron magnetic dipole points in the opposite direction from its spin angular momentum vector (as compared to the proton). This gives the neutron, in effect, a magnetic moment which resembles a negatively charged particle. This can be reconciled classically with a neutral neutron composed of a charge distribution in which the negative sub-parts of the neutron have a larger average radius of distribution, and therefore contribute more to the particle's magnetic dipole moment, than do the positive parts that are, on average, nearer the core.

Electric dipole moment

The Standard Model of particle physics predicts a tiny separation of positive and negative charge within the neutron leading to a permanent

beyond the Standard Model
generally lead to much larger predictions for the electric dipole moment of the neutron. Currently, there are at least four experiments trying to measure for the first time a finite neutron electric dipole moment, including:

Antineutron

The antineutron is the

CPT-symmetry puts strong constraints on the relative properties of particles and antiparticles, so studying antineutrons provides stringent tests on CPT-symmetry. The fractional difference in the masses of the neutron and antineutron is (9±6)×10−5. Since the difference is only about two standard deviations away from zero, this does not give any convincing evidence of CPT-violation.[14]

Neutron compounds

Dineutrons and tetraneutrons

The

dineutron is considered an unbound isotope with lifetimes around 10-22 seconds. The first evidence for this state was reported by Haddock et al. in 1965.[78]: 275  In 2012, Artemis Spyrou from Michigan State University and coworkers reported that they observed, for the first time, direct dineutron emission in the decay of 16Be. The dineutron character is evidenced by a small emission angle between the two neutrons. The authors measured the two-neutron separation energy to be 1.35(10) MeV, in good agreement with shell model calculations, using standard interactions for this mass region.[79]

Evidence for unbound clusters of 4 neutrons, or tetraneutron as resonances in the disintegration of beryllium-14 nuclei,[80] in 8He-8Be interactions,[81] and collisions of 4He nuclei give an estimated lifetime around 10-22 seconds.[82] These discoveries should deepen our understanding of the nuclear forces.[83][84]

Neutron stars and neutron matter

At extremely high pressures and temperatures, nucleons and electrons are believed to collapse into bulk neutronic matter, called

neutron matter. This is presumed to happen in neutron stars.[85]

The extreme pressure inside a neutron star may deform the neutrons into a cubic symmetry, allowing tighter packing of neutrons.[86]

Detection

The common means of detecting a charged particle by looking for a track of ionization (such as in a cloud chamber) does not work for neutrons directly. Neutrons that elastically scatter off atoms can create an ionization track that is detectable, but the experiments are not as simple to carry out; other means for detecting neutrons, consisting of allowing them to interact with atomic nuclei, are more commonly used. The commonly used methods to detect neutrons can therefore be categorized according to the nuclear processes relied upon, mainly neutron capture or elastic scattering.[87]

Neutron detection by neutron capture

A common method for detecting neutrons involves converting the energy released from neutron capture reactions into electrical signals. Certain nuclides have a high neutron capture cross section, which is the probability of absorbing a neutron. Upon neutron capture, the compound nucleus emits more easily detectable radiation, for example an alpha particle, which is then detected. The nuclides 3
He
, 6
Li
, 10
B
, 233
U
, 235
U
, 237
Np
, and 239
Pu
are useful for this purpose.

Neutron detection by elastic scattering

Neutrons can elastically scatter off nuclei, causing the struck nucleus to recoil. Kinematically, a neutron can transfer more energy to a light nucleus such as hydrogen or helium than to a heavier nucleus. Detectors relying on elastic scattering are called fast neutron detectors. Recoiling nuclei can ionize and excite further atoms through collisions. Charge and/or scintillation light produced in this way can be collected to produce a detected signal. A major challenge in fast neutron detection is discerning such signals from erroneous signals produced by gamma radiation in the same detector. Methods such as pulse shape discrimination can be used in distinguishing neutron signals from gamma-ray signals, although certain inorganic scintillator-based detectors have been developed [88][89] to selectively detect neutrons in mixed radiation fields inherently without any additional techniques.

Fast neutron detectors have the advantage of not requiring a moderator, and are therefore capable of measuring the neutron's energy, time of arrival, and in certain cases direction of incidence.

Sources and production

Free neutrons are unstable, although they have the longest half-life of any unstable subatomic particle by several orders of magnitude. Their half-life is still only about 10 minutes, so they can be obtained only from sources that produce them continuously.

Natural neutron background. A small natural background flux of free neutrons exists everywhere on Earth.[90] In the atmosphere and deep into the ocean, the "neutron background" is caused by muons produced by cosmic ray interaction with the atmosphere. These high-energy muons are capable of penetration to considerable depths in water and soil. There, in striking atomic nuclei, among other reactions they induce spallation reactions in which a neutron is liberated from the nucleus. Within the Earth's crust a second source is neutrons produced primarily by spontaneous fission of uranium and thorium present in crustal minerals. The neutron background is not strong enough to be a biological hazard, but it is of importance to very high resolution particle detectors that are looking for very rare events, such as (hypothesized) interactions that might be caused by particles of dark matter.[90] Recent research has shown that even thunderstorms can produce neutrons with energies of up to several tens of MeV.[91] Recent research has shown that the fluence of these neutrons lies between 10−9 and 10−13 per ms and per m2 depending on the detection altitude. The energy of most of these neutrons, even with initial energies of 20 MeV, decreases down to the keV range within 1 ms.[92]

Even stronger neutron background radiation is produced at the surface of Mars, where the atmosphere is thick enough to generate neutrons from cosmic ray muon production and neutron-spallation, but not thick enough to provide significant protection from the neutrons produced. These neutrons not only produce a Martian surface neutron radiation hazard from direct downward-going neutron radiation but may also produce a significant hazard from reflection of neutrons from the Martian surface, which will produce reflected neutron radiation penetrating upward into a Martian craft or habitat from the floor.[93]

Sources of neutrons for research. These include certain types of radioactive decay (spontaneous fission and neutron emission), and from certain nuclear reactions. Convenient nuclear reactions include tabletop reactions such as natural alpha and gamma bombardment of certain nuclides, often beryllium or deuterium, and induced nuclear fission, such as occurs in nuclear reactors. In addition, high-energy nuclear reactions (such as occur in cosmic radiation showers or accelerator collisions) also produce neutrons from disintegration of target nuclei. Small (tabletop) particle accelerators optimized to produce free neutrons in this way, are called neutron generators.

In practice, the most commonly used small laboratory sources of neutrons use radioactive decay to power neutron production. One noted neutron-producing

antimony-124 plus beryllium, a system with a half-life of 60.9 days, which can be constructed from natural antimony (which is 42.8% stable antimony-123) by activating it with neutrons in a nuclear reactor, then transported to where the neutron source is needed.[94]

Institut Laue–Langevin (ILL) in Grenoble, France – a major neutron research facility.

Nuclear fission reactors naturally produce free neutrons; their role is to sustain the energy-producing chain reaction. The intense neutron radiation can also be used to produce various radioisotopes through the process of neutron activation, which is a type of neutron capture.

Experimental

nuclear fusion-fission hybrids are proposed to make use of those neutrons to either maintain a subcritical reactor or to aid in nuclear transmutation
of harmful long lived nuclear waste to shorter lived or stable nuclides.

Neutron beams and modification of beams after production

Free neutron beams are obtained from neutron sources by neutron transport. For access to intense neutron sources, researchers must go to a specialized neutron facility that operates a research reactor or a spallation source.

The neutron's lack of total electric charge makes it difficult to steer or accelerate them. Charged particles can be accelerated, decelerated, or deflected by

Thermal neutrons can be polarized by transmission through magnetic materials in a method analogous to the Faraday effect for photons. Cold neutrons of wavelengths of 6–7 angstroms can be produced in beams of a high degree of polarization, by use of magnetic mirrors and magnetized interference filters.[97]

Applications

The neutron plays an important role in many nuclear reactions. For example, neutron capture often results in

radioactivity. In particular, knowledge of neutrons and their behavior has been important in the development of nuclear reactors and nuclear weapons. The fissioning of elements like uranium-235 and plutonium-239
is caused by their absorption of neutrons.

matter waves exhibit properties similar to geometrical and wave optics of light, including reflection, refraction, diffraction, and interference.[98] Neutrons are complementary to X-rays in terms of atomic contrasts by different scattering cross sections
; sensitivity to magnetism; energy range for inelastic neutron spectroscopy; and deep penetration into matter.

The development of "neutron lenses" based on total internal reflection within hollow glass capillary tubes or by reflection from dimpled aluminum plates has driven ongoing research into neutron microscopy and neutron/gamma ray tomography.[99][100][101][102]

A major use of neutrons is to excite delayed and prompt

bore holes
and industrial bulk materials on conveyor belts.

Another use of neutron emitters is the detection of light nuclei, in particular the hydrogen found in water molecules. When a fast neutron collides with a light nucleus, it loses a large fraction of its energy. By measuring the rate at which slow neutrons return to the probe after reflecting off of hydrogen nuclei, a neutron probe may determine the water content in soil.

Medical therapies

Because neutron radiation is both penetrating and ionizing, it can be exploited for medical treatments. However, neutron radiation can have the unfortunate side-effect of leaving the affected area radioactive. Neutron tomography is therefore not a viable medical application.

Fast neutron therapy uses high-energy neutrons typically greater than 20 MeV to treat cancer.

gamma radiation.[104]

Beams of low-energy neutrons are used in

that have sufficient energy to kill the malignant cell, but insufficient range to damage nearby cells. For such a therapy to be applied to the treatment of cancer, a neutron source having an intensity of the order of a thousand million (109) neutrons per second per cm2 is preferred. Such fluxes require a research nuclear reactor.

Protection

Exposure to free neutrons can be hazardous, since the interaction of neutrons with molecules in the body can cause disruption to molecules and atoms, and can also cause reactions that give rise to other forms of radiation (such as protons).[7] The normal precautions of radiation protection apply: Avoid exposure, stay as far from the source as possible, and keep exposure time to a minimum. But particular thought must be given to how to protect from neutron exposure. For other types of radiation, e.g., alpha particles, beta particles, or gamma rays, material of a high atomic number and with high density makes for good shielding; frequently, lead is used. However, this approach will not work with neutrons, since the absorption of neutrons does not increase straightforwardly with atomic number, as it does with alpha, beta, and gamma radiation. Instead one needs to look at the particular interactions neutrons have with matter (see the section on detection above). For example, hydrogen-rich materials are often used to shield against neutrons, since ordinary hydrogen both scatters and slows neutrons. This often means that simple concrete blocks or even paraffin-loaded plastic blocks afford better protection from neutrons than do far more dense materials. After slowing, neutrons may then be absorbed with an isotope that has high affinity for slow neutrons without causing secondary capture radiation, such as lithium-6.

Hydrogen-rich

CANDU-type reactors, in order to slow (moderate) neutron velocity, to increase the probability of nuclear fission compared to neutron capture
.

Neutron temperature

Thermal neutrons

Thermal neutrons are

free neutrons whose energies have a Maxwell–Boltzmann distribution with kT = 0.0253 eV (4.0×10−21 J) at room temperature. This gives characteristic (not average, or median) speed of 2.2 km/s. The name 'thermal' comes from their energy being that of the room temperature gas or material they are permeating. (see kinetic theory
for energies and speeds of molecules). After a number of collisions (often in the range of 10–20) with nuclei, neutrons arrive at this energy level, provided that they are not absorbed.

In many substances, thermal neutron reactions show a much larger effective cross-section than reactions involving faster neutrons, and thermal neutrons can therefore be absorbed more readily (i.e., with higher probability) by any

as a result.

Most

fast breeder
reactors, use fission energy neutrons directly.

Cold neutrons

Cold neutrons are thermal neutrons that have been equilibrated in a very cold substance such as liquid deuterium. Such a cold source is placed in the moderator of a research reactor or spallation source. Cold neutrons are particularly valuable for neutron scattering experiments.[105]

The use of cold and very cold neutrons (VCN) have been a bit limited compared to the use of thermal neutrons due to the relatively lower flux and lack in optical components. However, Innovative solutions have been proposed to offer more options to the scientific community to promote the use of VCN.[106][107]

Cold neutron source providing neutrons at about the temperature of liquid hydrogen

Ultracold neutrons

Ultracold neutrons are produced by inelastic scattering of cold neutrons in substances with a low neutron absorption cross section at a temperature of a few kelvins, such as solid deuterium[108] or superfluid helium.[109] An alternative production method is the mechanical deceleration of cold neutrons exploiting the Doppler shift.[110][111]

Fission energy neutrons

A fast neutron is a free neutron with a kinetic energy level close to MeV (1.6×10−13 J), hence a speed of ~14000 km/s (~ 5% of the speed of light). They are named fission energy or fast neutrons to distinguish them from lower-energy thermal neutrons, and high-energy neutrons produced in cosmic showers or accelerators. Fast neutrons are produced by nuclear processes such as nuclear fission. Neutrons produced in fission, as noted above, have a Maxwell–Boltzmann distribution of kinetic energies from 0 to ~14 MeV, a mean energy of 2 MeV (for 235U fission neutrons), and a mode of only 0.75 MeV, which means that more than half of them do not qualify as fast (and thus have almost no chance of initiating fission in fertile materials, such as 238U and 232Th).

Fast neutrons can be made into thermal neutrons via a process called moderation. This is done with a

light water, or graphite
are used to moderate neutrons.

Fusion neutrons

The fusion reaction rate increases rapidly with temperature until it maximizes and then gradually drops off. The D–T rate peaks at a lower temperature (about 70 keV, or 800 million kelvins) and at a higher value than other reactions commonly considered for fusion energy.

D–T (

MeV of kinetic energy and traveling at 17% of the speed of light
. D–T fusion is also the easiest fusion reaction to ignite, reaching near-peak rates even when the deuterium and tritium nuclei have only a thousandth as much kinetic energy as the 14.1 MeV that will be produced.

14.1 MeV neutrons have about 10 times as much energy as fission neutrons, and are very effective at fissioning even non-

heavy nuclei, and these high-energy fissions produce more neutrons on average than fissions by lower-energy neutrons. This makes D–T fusion neutron sources such as proposed tokamak power reactors useful for transmutation of transuranic waste. 14.1 MeV neutrons can also produce neutrons by knocking them loose from nuclei
.

On the other hand, these very high-energy neutrons are less likely to simply

reactor grade plutonium. This physical fact thus causes ordinary non-weapons grade materials to become of concern in certain nuclear proliferation
discussions and treaties.

Other fusion reactions produce much less energetic neutrons. D–D fusion produces a 2.45 MeV neutron and helium-3 half of the time, and produces tritium and a proton but no neutron the rest of the time. D–3He fusion produces no neutron.

Intermediate-energy neutrons

light water reactor
, which is a thermal-spectrum reactor

A fission energy neutron that has slowed down but not yet reached thermal energies is called an epithermal neutron.

transient changes in reactivity
that might make reactor control more difficult.

Ratios of capture reactions to fission reactions are also worse (more captures without fission) in most

thorium cycle
, which has good capture-fission ratios at all neutron energies.

High-energy neutrons

High-energy neutrons have much more energy than fission energy neutrons and are generated as secondary particles by particle accelerators or in the atmosphere from cosmic rays. These high-energy neutrons are extremely efficient at ionization and far more likely to cause cell death than X-rays or protons.[112][113]

See also

Neutron sources

Processes involving neutrons

References

  1. ^ Ernest Rutherford Archived 2011-08-03 at the Wayback Machine. Chemed.chem.purdue.edu. Retrieved on 2012-08-16.
  2. ^ a b 1935 Nobel Prize in Physics Archived 2017-10-03 at the Wayback Machine. Nobelprize.org. Retrieved on 2012-08-16.
  3. ^ a b "2018 CODATA recommended values" https://physics.nist.gov/cuu/Constants/index.html Archived 2018-01-22 at the Wayback Machine
  4. ^ a b c d e f Mohr, P.J.; Taylor, B.N. and Newell, D.B. (2014), "The 2014 CODATA Recommended Values of the Fundamental Physical Constants" Archived 2013-10-09 at the Wayback Machine (Web Version 7.0). The database was developed by J. Baker, M. Douma, and S. Kotochigova. (2014). National Institute of Standards and Technology, Gaithersburg, Maryland 20899.
  5. ^ Zyla, P. A. (2020). "n MEAN LIFE". PDG Live: 2020 Review of Particle Physics. Particle Data Group. Archived from the original on 17 January 2021. Retrieved 25 February 2021.
  6. ^ a b Olive, K.A.; (Particle Data Group); et al. (2014). "Review of Particle Physics" (PDF). (PDF) from the original on 2020-06-01. Retrieved 2017-10-26.
  7. ^
  8. .
  9. ^ Nudat 2 Archived 2009-08-17 at the Wayback Machine. Nndc.bnl.gov. Retrieved on 2010-12-04.
  10. ^ Particle Data Group Summary Data Table on Baryons Archived 2011-09-10 at the Wayback Machine. lbl.gov (2007). Retrieved on 2012-08-16.
  11. .
  12. ^ Basdevant, J.-L.; Rich, J.; Spiro, M. (2005). Fundamentals in Nuclear Physics. .
  13. ^ . The exact value of the mean lifetime is still uncertain, due to conflicting results from experiments. The Particle Data Group reports values up to six seconds apart (more than four standard deviations), commenting that "our 2006, 2008, and 2010 Reviews stayed with 885.7±0.8 s; but we noted that in light of SEREBROV 05 our value should be regarded as suspect until further experiments clarified matters. Since our 2010 Review, PICHLMAIER 10 has obtained a mean life of 880.7±1.8 s, closer to the value of SEREBROV 05 than to our average. And SEREBROV 10B[...] claims their values should be lowered by about 6 s, which would bring them into line with the two lower values. But those re-evaluations have not received an enthusiastic response from the experimenters in question; and in any case the Particle Data Group would have to await published changes (by those experimenters) of published values. At this point, we can think of nothing better to do than to average the seven best but discordant measurements, getting 881.5±1.5 s. Note that the error includes a scale factor of 2.7. This is a jump of 4.2 old (and 2.8 new) standard deviations. This state of affairs is a particularly unhappy one, because the value is so important. We again call upon the experimenters to clear this up."
  14. from the original on 2020-10-19. Retrieved 2020-10-14.
  15. ]
  16. ^ Sir James Chadwick's Discovery of Neutrons Archived 2011-10-26 at the Wayback Machine. ANS Nuclear Cafe. Retrieved on 2012-08-16.
  17. ^ Adair, R.K. (1989). The Great Design: Particles, Fields, and Creation. .
  18. ^ Cottingham, W.N.; Greenwood, D.A. (1986). An Introduction to Nuclear Physics. .
  19. ^ a b Rutherford, E. (1920). "Nuclear Constitution of Atoms". .
  20. .
  21. ^ .
  22. .
  23. .
  24. .
  25. ^ Friedlander G., Kennedy J.W. and Miller J.M. (1964) Nuclear and Radiochemistry (2nd edition), Wiley, pp. 22–23 and 38–39
  26. ^ .
  27. .
  28. .
  29. .
  30. .
  31. from the original on 2022-03-04. Retrieved 2012-06-16.
  32. .
  33. .
  34. ^ "Atop the Physics Wave: Rutherford Back in Cambridge, 1919–1937". Rutherford's Nuclear World. American Institute of Physics. 2011–2014. Archived from the original on 21 October 2014. Retrieved 19 August 2014.
  35. .
  36. .
  37. .
  38. .
  39. .
  40. , pp. 84–88.
  41. .
  42. .
  43. .
  44. .
  45. .
  46. .
  47. .
  48. .
  49. ^ "The Nobel Prize in Chemistry 1944". Nobel Foundation. Archived from the original on 2018-12-26. Retrieved 2007-12-17.
  50. .
  51. ^ "The Nobel Prize in Chemistry 1944: Presentation Speech". Nobel Foundation. Archived from the original on 2007-10-25. Retrieved 2008-01-03.
  52. PMID 10033294
    .
  53. , pp. 18–19
  54. ^ .
  55. (PDF) on 2016-08-13. Retrieved 2016-06-26.
  56. .
  57. .
  58. .
  59. from the original on 2022-04-07. Retrieved 2020-08-27.
  60. ^ .
  61. ^ .
  62. ^ Greenberg, O.W. (2009), "Color Charge Degree of Freedom in Particle Physics", Compendium of Quantum Physics, Springer Berlin Heidelberg, pp. 109–111,
    S2CID 17512393
  63. .
  64. .
  65. ^ a b Cho, Adrian (2 April 2010). "Mass of the Common Quark Finally Nailed Down". Science. American Association for the Advancement of Science. Archived from the original on 27 August 2015. Retrieved 27 September 2014.
  66. ^ Wilczek, F. (2003). "The Origin of Mass" (PDF). MIT Physics Annual: 24–35. Archived (PDF) from the original on June 20, 2015.
  67. ^ Ji, Xiangdong (1995). "A QCD Analysis of the Mass Structure of the Nucleon".
    S2CID 15148740
    .
  68. (PDF) from the original on 2020-04-20. Retrieved 2019-08-25.
  69. ^ Kincade, Kathy (2 February 2015). "Pinpointing the magnetic moments of nuclear matter". Phys.org. Archived from the original on 2 May 2015. Retrieved May 8, 2015.
  70. ^ Miller, G.A. (2007). "Charge Densities of the Neutron and Proton".
    S2CID 119120565
    .
  71. ^ "Pear-shaped particles probe big-bang mystery" (Press release). University of Sussex. 20 February 2006. Archived from the original on 2011-06-07. Retrieved 2009-12-14.
  72. ^ A cryogenic experiment to search for the EDM of the neutron Archived 2012-02-16 at the Wayback Machine. Hepwww.rl.ac.uk. Retrieved on 2012-08-16.
  73. ^ Search for the neutron electric dipole moment: nEDM Archived 2015-09-25 at the Wayback Machine. Nedm.web.psi.ch (2001-09-12). Retrieved on 2012-08-16.
  74. ^ US nEDM ORNL experiment public page Archived 2017-04-30 at the Wayback Machine. Retrieved on 2017-02-08.
  75. ^ SNS Neutron EDM Experiment Archived 2011-02-10 at the Wayback Machine. P25ext.lanl.gov. Retrieved on 2012-08-16.
  76. ^ Measurement of the Neutron Electric Dipole Moment Archived 2011-08-23 at the Wayback Machine. Nrd.pnpi.spb.ru. Retrieved on 2012-08-16.
  77. .
  78. .
  79. .
  80. .
  81. .
  82. ^ "Physicists find signs of four-neutron nucleus". 2016-02-24. Archived from the original on 2017-07-29. Retrieved 2017-06-27.
  83. .
  84. .
  85. .
  86. .
  87. .
  88. .
  89. ^ a b Carson, M.J.; et al. (2004). "Neutron background in large-scale xenon detectors for dark matter searches".
    S2CID 17887096
    .
  90. (PDF) from the original on 2019-12-23. Retrieved 2019-08-25.
  91. .
  92. PMID 11770546. Archived from the original
    (PDF) on 2005-02-25.
  93. , pp. 32–33.
  94. ^ "Isotopes and Radioactivity Tutorial". Archived from the original on 2020-02-14. Retrieved 2020-04-16.
  95. ^ Science/Nature |Q&A: Nuclear fusion reactor Archived 2022-02-25 at the Wayback Machine. BBC News (2006-02-06). Retrieved on 2010-12-04.
  96. , p. 453.
  97. .
  98. .
  99. ^ Physorg.com, "New Way of 'Seeing': A 'Neutron Microscope'" Archived 2012-01-24 at the Wayback Machine. Physorg.com (2004-07-30). Retrieved on 2012-08-16.
  100. ^ "NASA Develops a Nugget to Search for Life in Space" Archived 2014-03-08 at the Wayback Machine. NASA.gov (2007-11-30). Retrieved on 2012-08-16.
  101. ISSN 1044-8632
    .
  102. .
  103. ^ Johns HE and Cunningham JR (1978). The Physics of Radiology. Charles C Thomas 3rd edition
  104. ^ [email protected] (2017-04-17). "How neutrons are useful". NIST. Archived from the original on 2021-01-25. Retrieved 2021-01-21.
  105. .
  106. .
  107. .
  108. .
  109. .
  110. from the original on 2021-02-24. Retrieved 2020-04-24.
  111. ^ Freeman, Tami (May 23, 2008). "Facing up to secondary neutrons". Medical Physics Web. Archived from the original on 2010-12-20. Retrieved 2011-02-08.
  112. from the original on 2019-01-26. Retrieved 2019-01-25.

Further reading