Neutronium

Last updated

Neutronium (or neutrium, [1] or neutrite [2] ) is a hypothetical substance made purely of neutrons. The word was coined by scientist Andreas von Antropoff in 1926 (before the 1932 discovery of the neutron) for the hypothetical "element of atomic number zero" (with zero protons in its nucleus) that he placed at the head of the periodic table (denoted by -). [3] [4] However, the meaning of the term has changed over time, and from the last half of the 20th century onward it has been also used to refer to extremely dense substances resembling the neutron-degenerate matter theorized to exist in the cores of neutron stars; hereinafter "degenerate neutronium" will refer to this.

Contents

In neutron stars

Cross-section of neutron star. Here, the core has neutrons or neutron-degenerate matter and quark matter. Neutron star cross section.svg
Cross-section of neutron star. Here, the core has neutrons or neutron-degenerate matter and quark matter.

Neutronium is used in popular physics literature [1] [2] to refer to the material present in the cores of neutron stars (stars which are too massive to be supported by electron degeneracy pressure and which collapse into a denser phase of matter). In scientific literature the term "neutron-degenerate matter" [5] or simply neutron matter is used for this material. [6]

Hypothetical multi-neutrons

The term "neutronium" was coined in 1926 by Andreas von Antropoff for a conjectured form of matter made up of neutrons with no protons or electrons, which he placed as the chemical element of atomic number zero at the head of his new version of the periodic table. [3] It was subsequently placed in the middle of several spiral representations of the periodic system for classifying the chemical elements, such as those of Charles Janet (1928), Edgar Emerson (1944), [7] [8] and John D. Clark (1950).

The term is not used in the scientific literature either for a condensed form of matter, or as an element, and theoretical analysis expects no bound forms of neutrons without protons. [9] If neutronium were considered to be an element, then these neutron clusters could be considered to be the isotopes of that element. However, these reports have not been further substantiated.

See also

Related Research Articles

<span class="mw-page-title-main">Neutron</span> Subatomic particle with no charge

The neutron is a subatomic particle, symbol
n
or
n0
, which has a neutral 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. Their properties and interactions are described by nuclear physics. Protons and neutrons are not elementary particles; each is composed of three quarks.

<span class="mw-page-title-main">Proton</span> Subatomic particle with positive charge

A proton is a stable subatomic particle, symbol
p
, H+, or 1H+ with a positive electric charge of +1 e (elementary charge). Its mass is slightly less than the mass of a neutron and approximately 1836 times the mass of an electron (the proton-to-electron mass ratio). Protons and neutrons, each with a mass of approximately one atomic mass unit, are jointly referred to as nucleons (particles present in atomic nuclei).

<span class="mw-page-title-main">Big Bang nucleosynthesis</span> Process during the early universe

In physical cosmology, Big Bang nucleosynthesis is the production of nuclei other than those of the lightest isotope of hydrogen during the early phases of the universe. This type of nucleosynthesis is thought by most cosmologists to have occurred from 10 seconds to 20 minutes after the Big Bang. It is thought to be responsible for the formation of most of the universe's helium, along with small fractions of the hydrogen isotope deuterium, the helium isotope helium-3 (3He), and a very small fraction of the lithium isotope lithium-7 (7Li). In addition to these stable nuclei, two unstable or radioactive isotopes were produced: the heavy hydrogen isotope tritium and the beryllium isotope beryllium-7 (7Be). These unstable isotopes later decayed into 3He and 7Li, respectively, as above.

<span class="mw-page-title-main">Island of stability</span> Predicted set of isotopes of relatively more stable superheavy elements

In nuclear physics, the island of stability is a predicted set of isotopes of superheavy elements that may have considerably longer half-lives than known isotopes of these elements. It is predicted to appear as an "island" in the chart of nuclides, separated from known stable and long-lived primordial radionuclides. Its theoretical existence is attributed to stabilizing effects of predicted "magic numbers" of protons and neutrons in the superheavy mass region.

Tetraneutron is considered an unbound isotope with a lifetime around 10-22 seconds. The stability of this cluster of four neutrons is not supported by current models of nuclear forces. Recent empirical evidence is "consistent with a quasi-bound tetraneutron state existing for a very short time".

A hypernucleus is similar to a conventional atomic nucleus, but contains at least one hyperon in addition to the normal protons and neutrons. Hyperons are a category of baryon particles that carry non-zero strangeness quantum number, which is conserved by the strong and electromagnetic interactions.

<span class="mw-page-title-main">Halo nucleus</span> Core atomic nucleus surrounded by orbiting protons or neutrons

In nuclear physics, an atomic nucleus is called a halo nucleus or is said to have a nuclear halo when it has a core nucleus surrounded by a "halo" of orbiting protons or neutrons, which makes the radius of the nucleus appreciably larger than that predicted by the liquid drop model. Halo nuclei form at the extreme edges of the table of nuclides — the neutron drip line and proton drip line — and have short half-lives, measured in milliseconds. These nuclei are studied shortly after their formation in an ion beam.

Although there are nine known isotopes of helium (2He), only helium-3 and helium-4 are stable. All radioisotopes are short-lived, the longest-lived being 6
He
with a half-life of 806.92(24) milliseconds. The least stable is 10
He
, with a half-life of 260(40) yoctoseconds, although it is possible that 2
He
may have an even shorter half-life.

Flerovium (114Fl) is a synthetic element, and thus a standard atomic weight cannot be given. Like all synthetic elements, it has no stable isotopes. The first isotope to be synthesized was 289Fl in 1999. Flerovium has six known isotopes, along with the unconfirmed 290Fl, and possibly two nuclear isomers. The longest-lived isotope is 289Fl with a half-life of 1.9 seconds, but 290Fl may have a longer half-life of 19 seconds.

<span class="mw-page-title-main">Calcium-48</span> Isotope of calcium

Calcium-48 is a scarce isotope of calcium containing 20 protons and 28 neutrons. It makes up 0.187% of natural calcium by mole fraction. Although it is unusually neutron-rich for such a light nucleus, its beta decay is extremely hindered, and so the only radioactive decay pathway that it has been observed to undergo is the extremely rare process of double beta decay. Its half-life is about 6.4×1019 years, so for all practical purposes it can be treated as stable. One factor contributing to this unusual stability is that 20 and 28 are both magic numbers, making 48Ca a "doubly magic" nucleus.

<span class="mw-page-title-main">Nuclear matter</span> System of interacting nucleons

Nuclear matter is an idealized system of interacting nucleons that exists in several phases of exotic matter that, as of yet, are not fully established. It is not matter in an atomic nucleus, but a hypothetical substance consisting of a huge number of protons and neutrons held together by only nuclear forces and no Coulomb forces. Volume and the number of particles are infinite, but the ratio is finite. Infinite volume implies no surface effects and translational invariance.

In particle physics, hexaquarks, alternatively known as sexaquarks, are a large family of hypothetical particles, each particle consisting of six quarks or antiquarks of any flavours. Six constituent quarks in any of several combinations could yield a colour charge of zero; for example a hexaquark might contain either six quarks, resembling two baryons bound together, or three quarks and three antiquarks. Once formed, dibaryons are predicted to be fairly stable by the standards of particle physics.

<span class="mw-page-title-main">Quark–gluon plasma</span> Phase of quantum chromodynamics (QCD)

Quark–gluon plasma is an interacting localized assembly of quarks and gluons at thermal and chemical (abundance) equilibrium. The word plasma signals that free color charges are allowed. In a 1987 summary, Léon Van Hove pointed out the equivalence of the three terms: quark gluon plasma, quark matter and a new state of matter. Since the temperature is above the Hagedorn temperature—and thus above the scale of light u,d-quark mass—the pressure exhibits the relativistic Stefan-Boltzmann format governed by temperature to the fourth power and many practically massless quark and gluon constituents. It can be said that QGP emerges to be the new phase of strongly interacting matter which manifests its physical properties in terms of nearly free dynamics of practically massless gluons and quarks. Both quarks and gluons must be present in conditions near chemical (yield) equilibrium with their colour charge open for a new state of matter to be referred to as QGP.

A strangelet is a hypothetical particle consisting of a bound state of roughly equal numbers of up, down, and strange quarks. An equivalent description is that a strangelet is a small fragment of strange matter, small enough to be considered a particle. The size of an object composed of strange matter could, theoretically, range from a few femtometers across to arbitrarily large. Once the size becomes macroscopic, such an object is usually called a strange star. The term "strangelet" originates with Edward Farhi and Robert Jaffe in 1984. It has been theorized that strangelets can convert matter to strange matter on contact. Strangelets have also been suggested as a dark matter candidate.

The EMC effect is the surprising observation that the cross section for deep inelastic scattering from an atomic nucleus is different from that of the same number of free protons and neutrons. From this observation, it can be inferred that the quark momentum distributions in nucleons bound inside nuclei are different from those of free nucleons. This effect was first observed in 1983 at CERN by the European Muon Collaboration, hence the name "EMC effect". It was unexpected, since the average binding energy of protons and neutrons inside nuclei is insignificant when compared to the energy transferred in deep inelastic scattering reactions that probe quark distributions. While over 1000 scientific papers have been written on the topic and numerous hypotheses have been proposed, no definitive explanation for the cause of the effect has been confirmed. Determining the origin of the EMC effect is one of the major unsolved problems in the field of nuclear physics.

<span class="mw-page-title-main">Nuclear drip line</span> Atomic nuclei decay delimiter

The nuclear drip line is the boundary beyond which atomic nuclei are unbound with respect to the emission of a proton or neutron.

The rms charge radius is a measure of the size of an atomic nucleus, particularly the proton distribution. The proton radius is about one femtometre = 10−15 metre. It can be measured by the scattering of electrons by the nucleus. Relative changes in the mean squared nuclear charge distribution can be precisely measured with atomic spectroscopy.

<span class="mw-page-title-main">Borromean nucleus</span>

A Borromean nucleus is an atomic nucleus comprising three bound components in which any subsystem of two components is unbound. This has the consequence that if one component is removed, the remaining two comprise an unbound resonance, so that the original nucleus is split into three parts.

Reinhard Stock is a German experimental physicist, specializing in heavy-ion physics.

<span class="mw-page-title-main">Shape of the atomic nucleus</span>

The shape of the atomic nucleus depends on the variety of factors related to the size and shape of its nucleon constituents and the nuclear force holding them together. The origins of nuclear shape begin with the spacial extent of almost nearly all stable and a great many unstable nuclei has been determined mainly by electron and muon scattering experiments as well as spectroscopic experiments. An important factor in the internal structure of the nucleus is the nucleon-nucleon potential, which ultimately governs the distance between individual nucleons, while a dip in the charge density of some light nuclide structures a lesser density of nucleonic matter. A surprising non-spherical expectation for the shape of the nucleus originated in 1939 in the spectroscopic analysis of the quadrapole moments while the prolate spheroid shape of the nucleon arises from analysis of the intrinsic quadruple moment. The simple spherical approximation of nuclear size and shape provides at best a textbook introduction to nuclear size and shape. The unusual cosmic abundance of alpha nuclides has inspired geometric arrangements of alpha particles as a solution to nuclear shapes, although the atomic nucleus generally assumes a prolate spheroid shape. Nuclides can also be discus-shaped, triaxial or pear-shaped.

References

  1. 1 2 Inglis-Arkell, Esther (2012-04-14). "Neutrium: The Most Neutral Hypothetical State of Matter Ever". io9.com . Archived from the original on 2014-11-12. Retrieved 2013-02-11.
  2. 1 2 Zhuravleva, Valentina (2005). Ballad of the Stars: Stories of Science Fiction, Ultraimagination, and TRIZ. Technical Innovation Center, Inc. p. 75. ISBN   978-0-9640740-6-4. Archived from the original on 2022-04-12. Retrieved 2019-04-25.
  3. 1 2 von Antropoff, A. (1926). "Eine neue Form des periodischen Systems der Elementen". Zeitschrift für Angewandte Chemie (in German). 39 (23): 722–725. Bibcode:1926AngCh..39..722V. doi:10.1002/ange.19260392303.
  4. Stewart, P. J. (2007). "A century on from Dmitrii Mendeleev: Tables and spirals, noble gases and Nobel prizes". Foundations of Chemistry . 9 (3): 235–245. doi:10.1007/s10698-007-9038-x. S2CID   97131841.
  5. Angelo, J. A. (2006). Encyclopedia of Space and Astronomy. Infobase Publishing. p. 178. ISBN   978-0-8160-5330-8. Archived from the original on 2019-12-15. Retrieved 2016-10-28.
  6. Gandolfi, Stefano; Gezerlis, Alexandros; Carlson, J. (2015-10-19). "Neutron Matter from Low to High Density". Annual Review of Nuclear and Particle Science. 65 (1): 303–328. arXiv: 1501.05675 . Bibcode:2015ARNPS..65..303G. doi:10.1146/annurev-nucl-102014-021957. ISSN   0163-8998.
  7. Emerson, Edgar I. (1944). "A new spiral form of the periodic table". Journal of Chemical Education. 21 (3): 111. Bibcode:1944JChEd..21..111E. doi:10.1021/ed021p111.
  8. Emerson, Edgar I. (1944). "A chart based on atomic numbers showing the electronic structure of the elements". Journal of Chemical Education. 21 (5): 254. Bibcode:1944JChEd..21..254E. doi:10.1021/ed021p254.
  9. Timofeyuk, N. K. (2003). "Do multineutrons exist?". Journal of Physics G . 29 (2): L9. arXiv: nucl-th/0301020 . Bibcode:2003JPhG...29L...9T. doi:10.1088/0954-3899/29/2/102. S2CID   2847145.
  10. Schirber, M. (2012). "Nuclei Emit Paired-up Neutrons". Physics . 5: 30. Bibcode:2012PhyOJ...5...30S. doi:10.1103/Physics.5.30.
  11. Spyrou, A.; Kohley, Z.; Baumann, T.; Bazin, D.; et al. (2012). "First Observation of Ground State Dineutron Decay: 16Be". Physical Review Letters . 108 (10): 102501. Bibcode:2012PhRvL.108j2501S. doi: 10.1103/PhysRevLett.108.102501 . PMID   22463404.
  12. Bertulani, C. A.; Baur, G. (1986). "Coincidence Cross-sections for the Dissociation of Light Ions in High-energy Collisions" (PDF). Nuclear Physics A . 480 (3–4): 615–628. Bibcode:1988NuPhA.480..615B. doi:10.1016/0375-9474(88)90467-8. Archived from the original (PDF) on 2011-07-20.
  13. 1 2 Bertulani, C. A.; Canto, L. F.; Hussein, M. S. (1993). "The Structure And Reactions Of Neutron-Rich Nuclei" (PDF). Physics Reports . 226 (6): 281–376. Bibcode:1993PhR...226..281B. doi:10.1016/0370-1573(93)90128-Z. Archived from the original (PDF) on 2011-09-28.
  14. Hagino, K.; Sagawa, H.; Nakamura, T.; Shimoura, S. (2009). "Two-particle correlations in continuum dipole transitions in Borromean nuclei". Physical Review C . 80 (3): 1301. arXiv: 0904.4775 . Bibcode:2009PhRvC..80c1301H. doi:10.1103/PhysRevC.80.031301. S2CID   119293335.
  15. MacDonald, J.; Mullan, D. J. (2009). "Big Bang Nucleosynthesis: The Strong Nuclear Force meets the Weak Anthropic Principle". Physical Review D . 80 (4): 3507. arXiv: 0904.1807 . Bibcode:2009PhRvD..80d3507M. doi:10.1103/PhysRevD.80.043507. S2CID   119203730.
  16. Kneller, J. P.; McLaughlin, G. C. (2004). "The Effect of Bound Dineutrons upon BBN". Physical Review D . 70 (4): 3512. arXiv: astro-ph/0312388 . Bibcode:2004PhRvD..70d3512K. doi:10.1103/PhysRevD.70.043512. S2CID   119060865.
  17. Li, J. G.; Michel, N.; Hu, B. S.; Zuo, W.; Xu, F. R. (2019). "Ab initio no-core Gamow shell-model calculations of multineutron systems". Physical Review C. 100 (5): 054313. arXiv: 1911.06485 . Bibcode:2019PhRvC.100e4313L. doi:10.1103/PhysRevC.100.054313.
  18. Bertulani, C. A.; Zelevinsky, V. (2003). "Is the tetraneutron a bound dineutron-dineutron molecule?". Journal of Physics G . 29 (10): 2431–2437. arXiv: nucl-th/0212060 . Bibcode:2003JPhG...29.2431B. doi:10.1088/0954-3899/29/10/309. S2CID   55535943.
  19. "Tetra-Neutron Experiment: Understanding of Nuclear Forces Might Have To Be Significantly Changed". Archived 2021-12-13 at the Wayback Machine . SciTechDaily, December 12, 2021. Technical University of Munich (TUM)
  20. Bevelacqua, J. J. (1981). "Particle stability of the pentaneutron". Physics Letters B . 102 (2–3): 79–80. Bibcode:1981PhLB..102...79B. doi:10.1016/0370-2693(81)91033-9.