Unbihexium

Unbihexium, also known as element 126 or eka-plutonium, is the hypothetical chemical element with atomic number 126 and placeholder symbol Ubh. Unbihexium and Ubh are the temporary IUPAC name and symbol, respectively, until the element is discovered, confirmed, and a permanent name is decided upon. In the periodic table, unbihexium is expected to be a g-block superactinide and the eighth element in the 8th period. Unbihexium has attracted attention among nuclear physicists, especially in early predictions targeting properties of superheavy elements, for 126 may be a magic number of protons near the center of an island of stability, leading to longer half-lives, especially for 310Ubh or 354Ubh which may also have magic numbers of neutrons.[3]

Unbihexium, 126Ubh
Unbihexium
Pronunciation/ˌnbˈhɛksiəm/ (OON-by-HEKS-ee-em)
Alternative nameselement 126, eka-plutonium
Unbihexium in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson
Ununennium Unbinilium Unbiunium
Unquadquadium Unquadpentium Unquadhexium Unquadseptium Unquadoctium Unquadennium Unpentnilium Unpentunium Unpentbium Unpenttrium Unpentquadium Unpentpentium Unpenthexium Unpentseptium Unpentoctium Unpentennium Unhexnilium Unhexunium Unhexbium Unhextrium Unhexquadium Unhexpentium Unhexhexium Unhexseptium Unhexoctium Unhexennium Unseptnilium Unseptunium Unseptbium
Unbibium Unbitrium Unbiquadium Unbipentium Unbihexium Unbiseptium Unbioctium Unbiennium Untrinilium Untriunium Untribium Untritrium Untriquadium Untripentium Untrihexium Untriseptium Untrioctium Untriennium Unquadnilium Unquadunium Unquadbium Unquadtrium


Ubh

unbipentium ← unbihexium → unbiseptium
Atomic number (Z)126
Groupn/a
Periodperiod 8
Blockg-block
Element category  Unknown chemical properties, but probably a superactinide
Electron configuration[Og] 5g2 6f3 8s2 8p1 (predicted)[1]
Electrons per shell2, 8, 18, 32, 34, 21, 8, 3
(predicted)
Physical properties
unknown
Atomic properties
Oxidation states(+1), (+2), (+4), (+6), (+8) (predicted)[2]
Other properties
CAS Number54500-77-5
History
NamingIUPAC systematic element name

Early interest in possible increased stability led to the first attempted synthesis of unbihexium in 1971 and searches for it in nature in subsequent years. Despite several reported observations, more recent studies suggest that these experiments were insufficiently sensitive; hence, no unbihexium has been found naturally or artificially.[4] Predictions of the stability of unbihexium vary greatly among different models; some suggest the island of stability may instead lie at a lower atomic number, closer to copernicium and flerovium.

Unbihexium is predicted to be a chemically active superactinide, exhibiting a variety of oxidation states from +1 to +8, and possibly being a heavier congener of plutonium. It is predicted to be the second element with an electron in a g orbital, a consequence of relativistic effects seen only in heavy and superheavy elements. An overlap in energy levels of the 5g, 6f, 7d, and 8p orbitals is also expected, which complicates predictions of chemical properties for this element.

History

Synthesis attempts

The first and only attempt to synthesize unbihexium, which was unsuccessful, was performed in 1971 at CERN (European Organization for Nuclear Research) by René Bimbot and John M. Alexander using the hot fusion reaction:[3][5]

232
90
Th
+ 84
36
Kr
316
126
Ubh
* → no atoms

High-energy (13-15 MeV) alpha particles were observed and taken as possible evidence for the synthesis of unbihexium. Subsequent unsuccessful experiments with higher sensitivity suggest that the 10 mb sensitivity of this experiment was too low; hence, the formation of unbihexium nuclei in this reaction was deemed highly unlikely.[4]

Possible natural occurrence

A study in 1976 by a group of American researchers from several universities proposed that primordial superheavy elements, mainly livermorium, unbiquadium, unbihexium, and unbiseptium, with half-lives exceeding 500 million years[6] could be a cause of unexplained radiation damage (particularly radiohalos) in minerals.[4] This prompted many researchers to search for them in nature from 1976 to 1983. A group led by Tom Cahill, a professor at the University of California at Davis, claimed in 1976 that they had detected alpha particles and X-rays with the right energies to cause the damage observed, supporting the presence of these elements, especially unbihexium. Others claimed that none had been detected, and questioned the proposed characteristics of primordial superheavy nuclei.[4] In particular, they cited that the magic number N = 228 necessary for enhanced stability would create a neutron-excessive nucleus in unbihexium that might not be beta-stable, although several calculations suggest that 354Ubh may indeed be stable against beta decay.[7] This activity was also proposed to be caused by nuclear transmutations in natural cerium, raising further ambiguity upon this claimed observation of superheavy elements.[4]

Unbihexium has received particular attention in these investigations, for its speculated location in the island of stability may increase its abundance relative to other superheavy elements.[6] Any naturally occurring unbihexium is predicted to be chemically similar to plutonium and may exist with primordial 244Pu in the rare earth mineral bastnasite.[6] In particular, plutonium and unbihexium are predicted to have similar valence configurations, leading to the existence of unbihexium in the +4 oxidation state. Therefore, should unbihexium occur naturally, it may be possible to extract it using similar techniques for the accumulation of cerium and plutonium.[6] Likewise, unbihexium could also exist in monazite with other lanthanides and actinides that would be chemically similar.[4] Recent doubt on the existence of primordial 244Pu casts uncertainty on these predictions, however,[8] as the nonexistence (or minimal existence) of plutonium in bastnasite will inhibit possible identification of unbihexium as its heavier congener.

The possible extent of primordial superheavy elements on Earth today is uncertain. Even if they are confirmed to have caused the radiation damage long ago, they might now have decayed to mere traces, or even be completely gone.[9] It is also uncertain if such superheavy nuclei may be produced naturally at all, as spontaneous fission is expected to terminate the r-process responsible for heavy element formation between mass number 270 and 290, well before elements such as unbihexium may be formed.[10]

A recent hypothesis tries to explain the spectrum of Przybylski's Star by naturally occurring flerovium, unbinilium, and unbihexium.[11][12]

Naming

Using the 1979 IUPAC recommendations, the element should be temporarily called unbihexium (symbol Ubh) until it is discovered, the discovery is confirmed, and a permanent name chosen.[13] Although widely used in the chemical community on all levels, from chemistry classrooms to advanced textbooks, the recommendations are mostly ignored among scientists who work theoretically or experimentally on superheavy elements, who call it "element 126", with the symbol E126, (126), or 126.[14] Some researchers have also referred to unbihexium as eka-plutonium,[15][16] a name derived from the system Dmitri Mendeleev used to predict unknown elements, though such an extrapolation might not work for g-block elements with no known congeners, and eka-plutonium would instead refer to element 146[17] or 148[18] when the term is meant to denote the element directly below plutonium.

Difficulties in synthesis

Every element from mendelevium onward was produced in fusion-evaporation reactions, culminating in the discovery of the heaviest known element oganesson in 2002[19][20] and most recently tennessine in 2010.[21] These reactions approached the limit of current technology; for example, the synthesis of tennessine required 22 milligrams of 249Bk and an intense 48Ca beam for six months. The intensity of beams in superheavy element research cannot exceed 1012 projectiles per second without damaging the target and detector, and producing larger quantities of increasingly rare and unstable actinide targets is impractical.[22] Consequently, future experiments must be done at facilities such as the under-construction superheavy element factory (SHE-factory) at the Joint Institute for Nuclear Research (JINR) or RIKEN, which will allow experiments to run for longer stretches of time with increased detection capabilities and enable otherwise inaccessible reactions.[23] Even so, it will likely be a great challenge to synthesize elements beyond unbinilium (120) or unbiunium (121), given their short predicted half-lives and low predicted cross sections.[24]

It has been suggested that fusion-evaporation will not be feasible to reach unbihexium. As 48Ca cannot be used beyond elements 118 or possibly 119, the only alternatives are increasing the atomic number of the projectile or studying symmetric or near-symmetric reactions.[25] One calculation suggests that the cross section for producing unbihexium from 249Cf and 64Ni may be as low as nine orders of magnitude lower than the detection limit; such results are also suggested by the non-observation of unbinilium and unbibium in reactions with heavier projectiles and experimental cross section limits.[26] If Z = 126 represents a closed proton shell, compound nuclei may have greater survival probability and the use of 64Ni may be more feasible for producing nuclei with 122 < Z < 126, especially for compound nuclei near the closed shell at N = 184.[27] However, the cross section still might not exceed 1 fb, posing an obstacle that may only be overcome with more sensitive equipment.[28]

Predicted properties

Nuclear stability and isotopes

This nuclear chart used by the Japan Atomic Energy Agency predicts the decay modes of nuclei up to Z = 149 and N = 256. At Z = 126 (top right), the beta-stability line passes through a region of instability towards spontaneous fission (half-lives less than 1 nanosecond) and extends into a "cape" of stability near the N = 228 shell closure, where an island of stability centered at the possibly doubly magic isotope 354Ubh may exist.[29]
This diagram depicts shell gaps in the nuclear shell model. Shell gaps are created when more energy is required to reach the shell at the next higher energy level, thus resulting in a particularly stable configuration. For protons, the shell gap at Z = 82 corresponds to the peak of stability at lead, and while there is disagreement of the magicity of Z = 114 and Z = 120, a shell gap appears at Z = 126, thus suggesting that there may be a proton shell closure at unbihexium.[30]

Extensions of the nuclear shell model predicted that the next magic numbers after Z = 82 and N = 126 (corresponding to 208Pb, the heaviest stable nucleus) were Z = 126 and N = 184, making 310Ubh the next candidate for a doubly magic nucleus. These speculations led to interest in the stability of unbihexium as early as 1957; Gertrude Scharff Goldhaber was one of the first physicists to predict a region of increased stability in the vicinity of, and possibly centered at, unbihexium.[3] This notion of an "island of stability" comprising longer-lived superheavy nuclei was popularized by University of California professor Glenn Seaborg in the 1960s.[31]

In this region of the periodic table, N = 184 and N = 228 have been suggested as closed neutron shells,[32] and various atomic numbers, including Z = 126, have been proposed as closed proton shells.[lower-alpha 1] The extent of stabilizing effects in the region of unbihexium is uncertain, however, due to predictions of shifting or weakening of the proton shell closure and possible loss of double magicity.[32] More recent research predicts the island of stability to instead be centered at beta-stable isotopes of copernicium (291Cn and 293Cn)[25][33] or flerovium (Z = 114), which would place unbihexium well above the island and result in short half-lives regardless of shell effects.

Earlier models suggested the existence of long-lived nuclear isomers resistant to spontaneous fission in the region near 310Ubh, with half-lives on the order of millions or billions of years.[34] However, more rigorous calculations as early as the 1970s yielded contradictory results; it is now believed that the island of stability is not centered at 310Ubh, and thus will not enhance the stability of this nuclide. Instead, 310Ubh is thought to be very neutron-deficient and susceptible to alpha decay and spontaneous fission in less than a microsecond, and it may even lie at or beyond the proton drip line.[3][24][29] A 2016 calculation on the decay properties of 288-339Ubh upholds these predictions; the isotopes lighter than 313Ubh (including 310Ubh) may indeed lie beyond the drip line and decay by proton emission, 313-327Ubh will alpha decay, possibly reaching flerovium and livermorium isotopes, and heavier isotopes will decay by spontaneous fission.[35] This study and a quantum tunneling model predict alpha-decay half-lives under a microsecond for isotopes lighter than 318Ubh, rendering them impossible to identify experimentally.[35][36][lower-alpha 2] Hence, the isotopes 318-327Ubh may be synthesized and detected, and may even constitute a region of increased stability against fission around N ~ 198 with half-lives up to several seconds.[33]

Beyond this point, a "sea of instability" defined by very low fission barriers (caused by greatly increasing Coulomb repulsion in superheavy elements) and consequently fission half-lives on the order of 10−18 seconds are predicted among various models. Although the exact limit of stability for half-lives over one microsecond varies, stability against fission is strongly dependent on the N = 184 and N = 228 shell closures and rapidly drops off immediately beyond the influence of the shell closure.[24][29] Such an effect may be reduced, however, if nuclear deformation in intermediate isotopes may lead to a shift in magic numbers;[37] a similar phenomenon was observed in the deformed doubly magic nucleus 270Hs.[38] This shift could then lead to longer half-lives, perhaps on the order of days, for isotopes such as 342Ubh that would also lie on the beta-stability line.[37] A second island of stability for spherical nuclei may exist in unbihexium isotopes with many more neutrons, centered at 354Ubh and conferring additional stability in N = 228 isotones near the beta-stability line.[29] Originally, a short half-life of 39 milliseconds was predicted for 354Ubh toward spontaneous fission, though a partial alpha half-life for this isotope was predicted to be 18 years.[3] More recent analysis suggests that this isotope may have a half-life on the order of 100 years should the closed shells have strong stabilizing effects, placing it at the peak of an island of stability.[29] It may also be possible that 354Ubh is not doubly magic, as the Z = 126 shell is predicted to be relatively weak, or in some calculations, completely nonexistent. This suggests that any relative stability in unbihexium isotopes would be only due to neutron shell closures that may or may not have a stabilizing effect at Z = 126.[7][32]

Chemical

Unbihexium is expected to be the sixth member of a superactinide series and possibly a heavier congener of plutonium. In the superactinide series, the Aufbau principle is expected to break down due to relativistic effects, and an overlap of the energy levels of the 7d, 8p, and especially 5g and 6f orbitals is expected, which renders predictions of chemical and atomic properties of these elements very difficult.[39] The ground state electron configuration of unbihexium is thus predicted to be [Og] 5g2 6f3 8s2 8p1,[1] in contrast to [Og] 5g6 8s2 derived from Aufbau.

As with the other early superactinides, it is predicted that unbihexium will be able to lose all eight valence electrons in chemical reactions, rendering a variety of oxidation states up to +8 possible.[2] The +4 oxidation state is predicted to be most common, in addition to +2 and +6.[1][17] Unbihexium should be able to form the tetroxide UbhO4 and hexahalides UbhF6 and UbhCl6, the latter with a fairly strong bond dissociation energy of 2.68 eV.[40] It may also be possible for unbihexium to form a stable monofluoride UbhF. Calculations suggest that a diatomic UbhF molecule will feature a bond between the 5g orbital in unbihexium and the 2p orbital in fluorine, thus characterizing unbihexium as an element whose 5g electrons should actively participate in bonding.[15][16] It is also predicted that the Ubh6+ (in particular, in UbhF6) and Ubh7+ ions will have the electron configurations [Og] 5g2 and [Og] 5g1, respectively, in contrast to the [Og] 6f1 configuration seen in Ubt4+ and Ubq5+ that bears more resemblance to their actinide homologs.[2] The activity of 5g electrons may influence the chemistry of superactinides such as unbihexium in new ways that are difficult to predict, as no known elements have electrons in a g orbital in the ground state.[17]

See also

Notes

  1. Aromic numbers 114, 120, 122, 124 have also been suggested as closed proton shells in different models.
  2. While such nuclei may be synthesized and a series of decay signals may be registered, decays faster than one microsecond may pile up with subsequent signals and thus be indistinguishable, especially when multiple uncharacterized nuclei may be formed and emit a series of similar alpha particles. The main difficulty is thus attributing the decays to the correct parent nucleus, as a superheavy atom that decays before reaching the detector will not be registered at all.

References

  1. Hoffman, Darleane C.; Lee, Diana M.; Pershina, Valeria (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer Science+Business Media. ISBN 978-1-4020-3555-5.
  2. Pyykkö, Pekka (2011). "A suggested periodic table up to Z ≤ 172, based on Dirac–Fock calculations on atoms and ions". Physical Chemistry Chemical Physics. 13 (1): 161–8. Bibcode:2011PCCP...13..161P. doi:10.1039/c0cp01575j. PMID 20967377.
  3. Bemis, C.E.; Nix, J.R. (1977). "Superheavy elements - the quest in perspective" (PDF). Comments on Nuclear and Particle Physics. 7 (3): 65–78. ISSN 0010-2709.
  4. Hoffman, D.C; Ghiorso, A.; Seaborg, G.T. (2000). The Transuranium People: The Inside Story. Imperial College Press. ISBN 1-86094-087-0.
  5. Emsley, John (2011). Nature's Building Blocks: An A-Z Guide to the Elements (New ed.). New York, NY: Oxford University Press. p. 588. ISBN 978-0-19-960563-7.
  6. Sheline, R.K. (1976). "A Suggested Source of Element 126". Zeitschrift für Physik A. 279 (3): 255–257. Bibcode:1976ZPhyA.279..255S. doi:10.1007/BF01408296.
  7. Lodhi, M.A.K., ed. (March 1978). Superheavy Elements: Proceedings of the International Symposium on Superheavy Elements. Lubbock, Texas: Pergamon Press. ISBN 0-08-022946-8.
  8. Lachner, J.; et al. (2012). "Attempt to detect primordial 244Pu on Earth". Physical Review C. 85 (1): 015801. Bibcode:2012PhRvC..85a5801L. doi:10.1103/PhysRevC.85.015801.
  9. Emsley, John (2011). Nature's Building Blocks: An A–Z Guide to the Elements (New ed.). New York: Oxford University Press. p. 592. ISBN 978-0-19-960563-7.
  10. Petermann, I; Langanke, K.; Martínez-Pinedo, G.; Panov, I.V; Reinhard, P.G.; Thielemann, F.K. (2012). "Have superheavy elements been produced in nature?". European Physical Journal A. 48 (122). arXiv:1207.3432. Bibcode:2012EPJA...48..122P. doi:10.1140/epja/i2012-12122-6.
  11. Jason Wright (16 March 2017). "Przybylski's Star III: Neutron Stars, Unbinilium, and aliens". Retrieved 31 July 2018.
  12. V. A. Dzuba; V. V. Flambaum; J. K. Webb (2017). "Isotope shift and search for metastable superheavy elements in astrophysical data". Physical Review A. 95 (6): 062515. arXiv:1703.04250. Bibcode:2017PhRvA..95f2515D. doi:10.1103/PhysRevA.95.062515.
  13. Chatt, J. (1979). "Recommendations for the naming of elements of atomic numbers greater than 100". Pure and Applied Chemistry. 51 (2): 381–384. doi:10.1351/pac197951020381.
  14. Haire, Richard G. (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer Science+Business Media. p. 1724. ISBN 1-4020-3555-1.
  15. Malli, G.L. (2006). "Dissociation energy of ekaplutonium fluoride E126F: The first diatomic with molecular spinors consisting of g atomic spinors". The Journal of Chemical Physics. 124 (7): 071102. Bibcode:2006JChPh.124g1102M. doi:10.1063/1.2173233. PMID 16497023.
  16. Jacoby, Mitch (2006). "As-yet-unsynthesized superheavy atom should form a stable diatomic molecule with fluorine". Chemical & Engineering News. 84 (10): 19. doi:10.1021/cen-v084n010.p019a.
  17. Fricke, B.; Greiner, W.; Waber, J. T. (1971). "The continuation of the periodic table up to Z = 172. The chemistry of superheavy elements" (PDF). Theoretica Chimica Acta. 21 (3): 235–260. doi:10.1007/BF01172015.
  18. Nefedov, V.I.; Trzhaskovskaya, M.B.; Yarzhemskii, V.G. (2006). "Electronic Configurations and the Periodic Table for Superheavy Elements" (PDF). Doklady Physical Chemistry. 408 (2): 149–151. doi:10.1134/S0012501606060029. ISSN 0012-5016.
  19. Oganessian, YT; et al. (2002). "Element 118: results from the first 249
    Cf
    + 48
    Ca
    experiment"
    . Communication of the Joint Institute for Nuclear Research. Archived from the original on 22 July 2011.
  20. "Livermore scientists team with Russia to discover element 118". Livermore press release. 3 December 2006. Retrieved 18 January 2008.
  21. Oganessian, YT; Abdullin, F; Bailey, PD; et al. (2010). "Synthesis of a New Element with Atomic Number 117" (PDF). Physical Review Letters. 104 (142502): 142502. Bibcode:2010PhRvL.104n2502O. doi:10.1103/PhysRevLett.104.142502. PMID 20481935.
  22. Roberto, JB (2015). "Actinide Targets for Super-Heavy Element Research" (PDF). cyclotron.tamu.edu. Texas A & M University. Retrieved 30 October 2018.
  23. Hagino, Kouichi; Hofmann, Sigurd; Miyatake, Hiroari; Nakahara, Hiromichi (2012). "平成23年度 研究業績レビュー(中間レビュー)の実施について" (PDF). www.riken.jp. RIKEN. Retrieved 5 May 2017.
  24. Karpov, A; Zagrebaev, V; Greiner, W (2015). "Superheavy Nuclei: which regions of nuclear map are accessible in the nearest studies" (PDF). cyclotron.tamu.edu. Texas A & M University. Retrieved 30 October 2018.
  25. Zagrebaev, Valeriy; Karpov, Alexander; Greiner, Walter (2013). "Future of superheavy element research: Which nuclei could be synthesized within the next few years?" (PDF). Journal of Physics. 420: 012001. arXiv:1207.5700. Bibcode:2013JPhCS.420a2001Z. doi:10.1088/1742-6596/420/1/012001.
  26. Giardina, G.; Fazio, G.; Mandaglio, G.; Manganaro, M.; Nasirov, A.K.; Romaniuk, M.V.; Saccà, C. (2010). "Expectations and limits to synthesize nuclei with Z ≥ 120". International Journal of Modern Physics E. 19 (5 & 6): 882–893. Bibcode:2010IJMPE..19..882G. doi:10.1142/S0218301310015333.
  27. Rykaczewski, Krzysztof P. (July 2016). "Super Heavy Elements and Nuclei" (PDF). people.nscl.msu.edu. MSU. Retrieved 30 April 2017.
  28. Kuzmina, A.Z.; Adamian, G.G.; Antonenko, N.V.; Scheid, W. (2012). "Influence of proton shell closure on production and identification of new superheavy nuclei". Physical Review C. 85 (1): 014319. Bibcode:2012PhRvC..85a4319K. doi:10.1103/PhysRevC.85.014319.
  29. Koura, H. (2011). Decay modes and a limit of existence of nuclei in the superheavy mass region (PDF). 4th International Conference on the Chemistry and Physics of the Transactinide Elements. Retrieved 18 November 2018.
  30. Kratz, J. V. (5 September 2011). The Impact of Superheavy Elements on the Chemical and Physical Sciences (PDF). 4th International Conference on the Chemistry and Physics of the Transactinide Elements. Retrieved 27 August 2013.
  31. Considine, Glenn D.; Kulik, Peter H. (2002). Van Nostrand's scientific encyclopedia (9 ed.). Wiley-Interscience. ISBN 978-0-471-33230-5. OCLC 223349096.
  32. Koura, H.; Chiba, S. (2013). "Single-Particle Levels of Spherical Nuclei in the Superheavy and Extremely Superheavy Mass Region". Journal of the Physical Society of Japan. 82 (1): 014201. Bibcode:2013JPSJ...82a4201K. doi:10.7566/JPSJ.82.014201.
  33. Palenzuela, Y. M.; Ruiz, L. F.; Karpov, A.; Greiner, W. (2012). "Systematic Study of Decay Properties of Heaviest Elements" (PDF). Bulletin of the Russian Academy of Sciences: Physics. 76 (11): 1165–1171. Bibcode:2012BRASP..76.1165P. doi:10.3103/S1062873812110172. ISSN 1062-8738.
  34. Maly, J.; Walz, D.R. (1980). "Search for superheavy elements among fossil fission tracks in zircon" (PDF).
  35. Santhosh, K.P.; Priyanka, B.; Nithya, C. (2016). "Feasibility of observing the α decay chains from isotopes of SHN with Z = 128, Z = 126, Z = 124 and Z = 122". Nuclear Physics A. 955 (November 2016): 156–180. arXiv:1609.05498. Bibcode:2016NuPhA.955..156S. doi:10.1016/j.nuclphysa.2016.06.010.
  36. Chowdhury, R. P.; Samanta, C.; Basu, D.N. (2008). "Nuclear half-lives for α -radioactivity of elements with 100 ≤ Z ≤ 130". Atomic Data and Nuclear Data Tables. 94 (6): 781–806. arXiv:0802.4161. Bibcode:2008ADNDT..94..781C. doi:10.1016/j.adt.2008.01.003.
  37. Okunev, V.S. (2018). "About islands of stability and limiting mass of the atomic nuclei". IOP Conference Series: Materials Science and Engineering. 468: 012012–1–012012–13. doi:10.1088/1757-899X/468/1/012012.
  38. Dvorak, J.; et al. (2006). "Doubly Magic Nucleus 270
    108
    Hs
    162
    "
    . Physical Review Letters. 97 (24): 242501. Bibcode:2006PhRvL..97x2501D. doi:10.1103/PhysRevLett.97.242501. PMID 17280272.
  39. Seaborg (c. 2006). "transuranium element (chemical element)". Encyclopædia Britannica. Retrieved 2010-03-16.
  40. Malli, G.L. (2007). "Thirty years of relativistic self-consistent field theory for molecules: relativistic and electron correlation effects for atomic and molecular systems of transactinide superheavy elements up to ekaplutonium E126 with g-atomic spinors in the ground state configuration". Theoretical Chemistry Accounts. 118 (3): 473–482. doi:10.1007/s00214-007-0335-1.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.