Copernicium

Copernicium,  112Cn
General properties
Pronunciation /ˌkpərˈnɪsiəm/ (KOH-pər-NIS-ee-əm)
Mass number 285 (most stable isotope)
Copernicium 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
Hg

Cn

(Uhh)
roentgeniumcoperniciumnihonium
Atomic number (Z) 112
Group group 12
Period period 7
Element category   post-transition metal, alternatively considered a transition metal
Block d-block
Electron configuration [Rn] 5f14 6d10 7s2 (predicted)[1]
Electrons per shell
2, 8, 18, 32, 32, 18, 2 (predicted)
Physical properties
Phase at STP gas[2] (predicted)
Boiling point 357+112
−108
 K (84+112
−108
 °C, 183+202
−194
 °F)[3]
Density when liquid (at m.p.) 23.7 g/cm3 (predicted)[1]
Atomic properties
Oxidation states (4), 2, (1), 0 (parenthesized oxidation states are predictions)[1][4][5]
Ionization energies
  • 1st: 1155 kJ/mol
  • 2nd: 2170 kJ/mol
  • 3rd: 3160 kJ/mol
  • (more) (all estimated)[1]
Atomic radius calculated: 147 pm[1][5] (predicted)
Covalent radius 122 pm (predicted)[6]
Other properties
Crystal structure body-centered cubic (bcc)
Body-centered cubic crystal structure for copernicium

(predicted)[7]
CAS Number 54084-26-3
History
Naming after Nicolaus Copernicus
Discovery Gesellschaft für Schwerionenforschung (1996)
Main isotopes of copernicium
Iso­tope Abun­dance Half-life (t1/2) Decay mode Pro­duct
286Cn syn 8.45 s ? SF
285Cn syn 28 s α 281Ds
284Cn syn 98 ms SF
283Cn syn 4.2 s[8] 90% α 279Ds
10% SF
EC? 283Rg
282Cn syn 0.91 ms SF
281Cn syn 0.13 s SF 277Ds
277Cn syn 0.69 ms SF 273Ds

Copernicium is a synthetic chemical element with symbol Cn and atomic number 112. It is an extremely radioactive element, and can only be created in a laboratory. The most stable known isotope, copernicium-285, has a half-life of approximately 29 seconds. Copernicium was first created in 1996 by the GSI Helmholtz Centre for Heavy Ion Research near Darmstadt, Germany. It is named after the astronomer Nicolaus Copernicus.

In the periodic table of the elements, copernicium is a d-block transactinide element. During reactions with gold, it has been shown[9] to be an extremely volatile metal and a group 12 element, so much so that it is probably a gas at standard temperature and pressure.

Copernicium is calculated to have several properties that differ from its lighter homologues in group 12, zinc, cadmium and mercury; due to relativistic effects, it may even give up its 6d electrons instead of its 7s ones. Copernicium has also been calculated to possibly show the oxidation state +4, while mercury shows it in only one compound of disputed existence and zinc and cadmium do not show it at all. It has also been predicted to be more difficult to oxidize copernicium from its neutral state than the other group 12 elements.

History

Discovery

Copernicium was first created on February 9, 1996, at the Gesellschaft für Schwerionenforschung (GSI) in Darmstadt, Germany, by Sigurd Hofmann, Victor Ninov et al.[10] This element was created by firing accelerated zinc-70 nuclei at a target made of lead-208 nuclei in a heavy ion accelerator. A single atom (a second was reported but was found to have been based on data fabricated by Ninov) of copernicium was produced with a mass number of 277.[10]

208
82
Pb + 70
30
Zn → 278
112
Cn* → 277
112
Cn + 1
0
n

In May 2000, the GSI successfully repeated the experiment to synthesize a further atom of copernicium-277.[11][12] This reaction was repeated at RIKEN using the Search for a Super-Heavy Element Using a Gas-Filled Recoil Separator set-up in 2004 and 2013 to synthesize three further atoms and confirm the decay data reported by the GSI team.[13][14] This reaction had also previously been tried in 1971 at the Joint Institute for Nuclear Research in Dubna, Russia to aim for 276Cn (produced in the 2n channel), but without success.[15]

The IUPAC/IUPAP Joint Working Party (JWP) assessed the claim of copernicium's discovery by the GSI team in 2001[16] and 2003.[17] In both cases, they found that there was insufficient evidence to support their claim. This was primarily related to the contradicting decay data for the known nuclide rutherfordium-261. However, between 2001 and 2005, the GSI team studied the reaction 248Cm(26Mg,5n)269Hs, and were able to confirm the decay data for hassium-269 and rutherfordium-261. It was found that the existing data on rutherfordium-261 was for an isomer,[18] now designated rutherfordium-261m.

In May 2009, the JWP reported on the claims of discovery of element 112 again and officially recognized the GSI team as the discoverers of element 112.[19] This decision was based on the confirmation of the decay properties of daughter nuclei as well as the confirmatory experiments at RIKEN.[20]

Work had also been done at the Joint Institute for Nuclear Research in Dubna, Russia from 1998 to synthesise the heavier isotope 283Cn in the hot fusion reaction 238U(48Ca,3n)283Cn; most observed atoms of 283Cn decayed by spontaneous fission, although an alpha decay branch to 279Ds was detected. While initial experiments aimed to assign the produced nuclide with its observed long half-life of 3 minutes based on its chemical behaviour, this was found to be not mercury-like as would have been expected (copernicium being under mercury in the periodic table),[20] and indeed now it appears that the long-lived activity might not have been from 283Cn at all, but its electron capture daughter 283Rg instead, with a shorter-lived 4-second half-life associated with 283Cn. (Another possibility is assignment to a metastable isomeic state, 283mCn.)[21] While later cross-bombardments in the 242Pu+48Ca and 245Cm+48Ca reactions succeeded in confirming the properties of 283Cn and its parents 287Fl and 291Lv, and played a major role in the acceptance of the discoveries of flerovium and livermorium (elements 114 and 116) by the JWP in 2011, this work originated subsequent to the GSI's work on 277Cn and priority was assigned to the GSI.[20]

Naming

a painted portrait of Copernicus
Nicolaus Copernicus, who formulated a heliocentric model with the planets orbiting around the Sun, replacing Ptolemy's earlier geocentric model.

Using Mendeleev's nomenclature for unnamed and undiscovered elements, copernicium should be known as eka-mercury. In 1979, IUPAC published recommendations according to which the element was to be called ununbium (with the corresponding symbol of Uub),[22] a systematic element name as a placeholder, until the element was discovered (and the discovery then confirmed) and a permanent name was decided on. Although widely used in the chemical community on all levels, from chemistry classrooms to advanced textbooks, the recommendations were mostly ignored among scientists in the field, who either called it "element 112", with the symbol of E112, (112), or even simply 112.[1]

After acknowledging the GSI team's discovery, the IUPAC asked them to suggest a permanent name for element 112.[20][23] On 14 July 2009, they proposed copernicium with the element symbol Cp, after Nicolaus Copernicus "to honor an outstanding scientist, who changed our view of the world".[24]

During the standard six-month discussion period among the scientific community about the naming,[25][26] it was pointed out that the symbol Cp was previously associated with the name cassiopeium (cassiopium), now known as lutetium (Lu).[27][28] For this reason, the IUPAC disallowed the use of Cp as a future symbol, prompting the GSI team to put forward the symbol Cn as an alternative. On 19 February 2010, the 537th anniversary of Copernicus' birth, IUPAC officially accepted the proposed name and symbol.[25][29]

Isotopes

List of copernicium isotopes
Isotope
Half-life
[30]
Decay
mode[30]
Discovery
year
Reaction
277Cn0.69 msα1996208Pb(70Zn,n)
278Cn10? msα, SF ?unknown
279Cn0.2? ms[31]α, SF ?unknown
280Cn0.5? ms[31]α, SF ?unknown
281Cn97 msα2010285Fl(—,α)
282Cn0.8 msSF2002294Og(—,3α)
283Cn4 sα, SF, EC?1998238U(48Ca,3n)
284Cn97 msα, SF2002288Fl(—,α)
285Cn29 sα1999289Fl(—,α)
286Cn8.45 s ?SF2016294Lv(—,2α)

Copernicium has no stable or naturally occurring isotopes. Several radioactive isotopes have been synthesized in the laboratory, either by fusing two atoms or by observing the decay of heavier elements. Seven different isotopes have been reported with atomic masses from 281 to 286, and 277. Most of these decay predominantly through alpha decay, but some undergo spontaneous fission, and copernicium-283 may have an electron capture branch.[30]

The isotope copernicium-283 was instrumental in the confirmation of the discoveries of the elements flerovium and livermorium.[32]

Half-lives

All copernicium isotopes are extremely unstable and radioactive; in general, heavier isotopes are more stable than the lighter. The most stable isotope, 285Cn, has a half-life of 29 seconds; 283Cn has a half-life of 4 seconds, and the unconfirmed 286Cn has a half-life of about 8.45 seconds. Other isotopes have half-lives shorter than 0.1 seconds. 281Cn and 284Cn both have half-lives of 97 ms, and the other two isotopes have half-lives slightly under one millisecond.[30] It is predicted that the heavy isotopes 291Cn and 293Cn may have half-lives longer than a few decades, and may have been produced in the r-process and be detectable in cosmic rays, though they would be about 10−12 times as abundant as lead.[33]

The lightest isotopes of copernicium have been synthesized by direct fusion between two lighter nuclei and as decay products (except for 277Cn, which is not known to be a decay product), while the heavier isotopes are only known to be produced by decay of heavier nuclei. The heaviest isotope produced by direct fusion is 283Cn; the three heavier isotopes, 284Cn, 285Cn, and 286Cn, have only been observed as decay products of elements with larger atomic numbers.[30]

In 1999, American scientists at the University of California, Berkeley, announced that they had succeeded in synthesizing three atoms of 293Og.[34] These parent nuclei were reported to have successively emitted three alpha particles to form copernicium-281 nuclei, which were claimed to have undergone an alpha decay, emitting an alpha particle with decay energy of 10.68 MeV and half-life 0.90 ms, but their claim was retracted in 2001.[35] The isotope, however, was produced in 2010 by the same team. The new data contradicted the previous (fabricated)[36] data.[37]

Predicted properties

Chemical

Copernicium is the tenth and last member of the 6d series and is the heaviest group 12 element in the periodic table, below zinc, cadmium and mercury. It is predicted to differ significantly from the lighter group 12 elements. The valence s-subshells of the group 12 elements and period 7 elements are expected to be relativistically contracted most strongly at copernicium. This and the closed-shell configuration of copernicium result in it probably being a very noble metal. Its metallic bonds should also be very weak, possibly making it extremely volatile, like the noble gases, and potentially making it gaseous at room temperature.[1][38] However, it should be able to form metal–metal bonds with copper, palladium, platinum, silver, and gold; these bonds are predicted to be only about 15–20 kJ/mol weaker than the analogous bonds with mercury.[1]

Once copernicium is ionized, its chemistry may present several differences from those of zinc, cadmium, and mercury. Due to the stabilization of 7s electronic orbitals and destabilization of 6d ones caused by relativistic effects, Cn2+ is likely to have a [Rn]5f146d87s2 electronic configuration, using the 6d orbitals before the 7s one, unlike its homologues. The fact that the 6d electrons participate more readily in chemical bonding means that once copernicium is ionized, it may behave more like a transition metal than its lighter homologues, especially in the possible +4 oxidation state. In aqueous solutions, copernicium may form the +2 and perhaps +4 oxidation states.[1] The diatomic ion Hg2+
2
, featuring mercury in the +1 oxidation state, is well-known, but the Cn2+
2
ion is predicted to be unstable or even non-existent.[1] Copernicium(II) fluoride, CnF2, should be more unstable than the analogous mercury compound, mercury(II) fluoride (HgF2), and may even decompose spontaneously into its constituent elements. In polar solvents, copernicium is predicted to preferentially form the CnF
5
and CnF
3
anions rather than the analogous neutral fluorides (CnF4 and CnF2, respectively), although the analogous bromide or iodide ions may be more stable towards hydrolysis in aqueous solution. The anions CnCl2−
4
and CnBr2−
4
should also be able to exist in aqueous solution.[1] Nevertheless, more recent experiments have cast doubt on the possible existence of HgF4, and indeed some calculations suggest that both HgF4 and CnF4 are actually unbound and of doubtful existence.[39] Analogous to mercury(II) cyanide (Hg(CN)2), copernicium is expected to form a stable cyanide, Cn(CN)2.[40]

Physical and atomic

Copernicium should be a very heavy metal with a density of around 23.7 g/cm3 in the solid state; in comparison, the most dense known element that has had its density measured, osmium, has a density of only 22.61 g/cm3. This results from copernicium's high atomic weight, the lanthanide and actinide contractions, and relativistic effects, although production of enough copernicium to measure this quantity would be impractical, and the sample would quickly decay.[1] However, some calculations predict copernicium to be a gas at room temperature, the first gaseous metal in the periodic table[1][38] (the second being flerovium, eka-lead), due to the closed-shell electron configurations of copernicium and flerovium.[41] The atomic radius of copernicium is expected to be around 147 pm. Due to the relativistic stabilization of the 7s orbital and destabilization of the 6d orbital, the Cn+ and Cn2+ ions are predicted to give up 6d electrons instead of 7s electrons, which is the opposite of the behavior of its lighter homologues.[1]

In addition to the relativistic contraction and binding of the 7s subshell, the 6d5/2 orbital is expected to be destabilized due to spin-orbit coupling, making it behave similarly to the 7s orbital in terms of size, shape, and energy. Calculations in 2007 expected that copernicium may therefore be a semiconductor[3] with a band gap of around 0.2 eV,[42] crystallizing in the hexagonal close-packed crystal structure.[42] However, calculations in 2017 and 2018 suggested that copernicium should be a noble metal at standard conditions with a body-centered cubic crystal structure: it should hence have no band gap, like mercury, although the density of states at the Fermi level is expected to be lower for copernicium than for mercury.[7][43] Like mercury, radon, and flerovium, but not oganesson (eka-radon), copernicium is calculated to have no electron affinity.[44]

Experimental atomic gas phase chemistry

Interest in copernicium's chemistry was sparked by predictions that it would have the largest relativistic effects in the whole of period 7 and group 12, and indeed among all 118 known elements.[1] Copernicium has the ground state electron configuration [Rn]5f146d107s2 and thus should belong to group 12 of the periodic table, according to the Aufbau principle. As such, it should behave as the heavier homologue of mercury and form strong binary compounds with noble metals like gold. Experiments probing the reactivity of copernicium have focused on the adsorption of atoms of element 112 onto a gold surface held at varying temperatures, in order to calculate an adsorption enthalpy. Owing to relativistic stabilization of the 7s electrons, copernicium shows radon-like properties. Experiments were performed with the simultaneous formation of mercury and radon radioisotopes, allowing a comparison of adsorption characteristics.[45]

The first chemical experiments on copernicium were conducted using the 238U(48Ca,3n)283Cn reaction. Detection was by spontaneous fission of the claimed parent isotope with half-life of 5 minutes. Analysis of the data indicated that copernicium was more volatile than mercury and had noble gas properties. However, the confusion regarding the synthesis of copernicium-283 has cast some doubt on these experimental results.[45] Given this uncertainty, between April–May 2006 at the JINR, a FLNR–PSI team conducted experiments probing the synthesis of this isotope as a daughter in the nuclear reaction 242Pu(48Ca,3n)287Fl.[45] (The 242Pu + 48Ca fusion reaction has a slightly larger cross-section than the 238U + 48Ca reaction, so that the best way to produce copernicium for chemical experimentation is as an overshoot product as the daughter of flerovium.)[46] In this experiment, two atoms of copernicium-283 were unambiguously identified and the adsorption properties indicated that copernicium is a more volatile homologue of mercury, due to formation of a weak metal-metal bond with gold, placing it firmly in group 12.[45] This agrees with general indications from relativistic calculations that copernicium is "more or less" homologous to mercury.[47]

In April 2007, this experiment was repeated and a further three atoms of copernicium-283 were positively identified. The adsorption property was confirmed and indicated that copernicium has adsorption properties completely in agreement with being the heaviest member of group 12.[45] These experiments also allowed the first experimental estimation of copernicium's boiling point: 84+112
−108
 °C, so that it may be a gas at standard conditions.[3]

Because the lighter group 12 elements often occur as chalcogenide ores, experiments were conducted in 2015 to attempt to deposit copernicium atoms on a selenium surface to form copernicium selenide, CnSe. Reaction of copernicium atoms with trigonal selenium to form a selenide were observed, with ΔHadsCn(t-Se) > 48 kJ/mol, with the kinetic hindrance towards selenide formation being lower for copernicium than for mercury. This was unexpected as the stability of the group 12 selenides tends to decrease down the group from ZnSe to HgSe, while it increases down the group for the group 14 selenides from GeSe to PbSe.[48]

See also

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Hoffman, Darleane C.; Lee, Diana M.; Pershina, Valeria (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean. The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer Science+Business Media. ISBN 1-4020-3555-1.
  2. Soverna S 2004, 'Indication for a gaseous element 112,' in U Grundinger (ed.), GSI Scientific Report 2003, GSI Report 2004-1, p. 187, ISSN 0174-0814
  3. 1 2 3 Eichler, R.; Aksenov, N. V.; Belozerov, A. V.; Bozhikov, G. A.; Chepigin, V. I.; Dmitriev, S. N.; Dressler, R.; Gäggeler, H. W.; et al. (2008). "Thermochemical and physical properties of element 112". Angewandte Chemie. 47 (17): 3262–6. doi:10.1002/anie.200705019. Retrieved 5 November 2013.
  4. Gäggeler, Heinz W.; Türler, Andreas (2013). "Gas Phase Chemistry of Superheavy Elements". The Chemistry of Superheavy Elements. Springer Science+Business Media. pp. 415–483. Retrieved 2018-04-21.
  5. 1 2 Fricke, Burkhard (1975). "Superheavy elements: a prediction of their chemical and physical properties". Recent Impact of Physics on Inorganic Chemistry. 21: 89–144. doi:10.1007/BFb0116498. Retrieved 4 October 2013.
  6. Chemical Data. Copernicium - Cn, Royal Chemical Society
  7. 1 2 Gyanchandani, Jyoti; Mishra, Vinayak; Dey, G. K.; Sikka, S. K. (January 2018). "Super heavy element Copernicium: Cohesive and electronic properties revisited". Solid State Communications. 269: 16–22. doi:10.1016/j.ssc.2017.10.009. Retrieved 28 March 2018.
  8. Chart of Nuclides. Brookhaven National Laboratory
  9. Eichler, R.; et al. (2007). "Chemical Characterization of Element 112". Nature. 447 (7140): 72–75. Bibcode:2007Natur.447...72E. doi:10.1038/nature05761. PMID 17476264.
  10. 1 2 Hofmann, S.; et al. (1996). "The new element 112". Zeitschrift für Physik A. 354 (1): 229–230. doi:10.1007/BF02769517.
  11. Hofmann, S.; et al. (2002). "New Results on Element 111 and 112". European Physical Journal A. 14 (2): 147–57. doi:10.1140/epja/i2001-10119-x.
  12. Hofmann, S.; et al. (2000). "New Results on Element 111 and 112" (PDF). Gesellschaft für Schwerionenforschung. Archived from the original (PDF) on February 27, 2008. Retrieved March 2, 2008.
  13. Morita, K. (2004). "Decay of an Isotope 277112 produced by 208Pb + 70Zn reaction". In Penionzhkevich, Yu. E.; Cherepanov, E. A. Exotic Nuclei: Proceedings of the International Symposium. World Scientific. pp. 188–191. doi:10.1142/9789812701749_0027.
  14. Popeko, Andrey G. (2016). "Synthesis of superheavy elements" (PDF). jinr.ru. Joint Institute for Nuclear Research. Retrieved 4 February 2018.
  15. Karol, P. J.; Nakahara, H.; Petley, B. W.; Vogt, E. (2001). "On the Discovery of the Elements 110–112" (PDF). Pure and Applied Chemistry. 73 (6): 959–967. doi:10.1351/pac200173060959.
  16. Karol, P. J.; Nakahara, H.; Petley, B. W.; Vogt, E. (2003). "On the Claims for Discovery of Elements 110, 111, 112, 114, 116 and 118" (PDF). Pure and Applied Chemistry. 75 (10): 1061–1611. doi:10.1351/pac200375101601.
  17. Dressler, R.; Türler, A. (2001). "Evidence for Isomeric States in 261Rf" (PDF). Annual Report. Paul Scherrer Institute. Archived from the original (PDF) on 2011-07-07.
  18. "A New Chemical Element in the Periodic Table". Gesellschaft für Schwerionenforschung. June 10, 2009. Archived from the original on August 23, 2009. Retrieved April 14, 2012.
  19. 1 2 3 4 Barber, R. C.; et al. (2009). "Discovery of the element with atomic number 112" (PDF). Pure and Applied Chemistry. 81 (7): 1331. doi:10.1351/PAC-REP-08-03-05.
  20. Hofmann, S.; Heinz, S.; Mann, R.; Maurer, J.; Münzenberg, G.; Antalic, S.; Barth, W.; Burkhard, H. G.; Dahl, L.; Eberhardt, K.; Grzywacz, R.; Hamilton, J. H.; Henderson, R. A.; Kenneally, J. M.; Kindler, B.; Kojouharov, I.; Lang, R.; Lommel, B.; Miernik, K.; Miller, D.; Moody, K. J.; Morita, K.; Nishio, K.; Popeko, A. G.; Roberto, J. B.; Runke, J.; Rykaczewski, K. P.; Saro, S.; Schneidenberger, C.; Schött, H. J.; Shaughnessy, D. A.; Stoyer, M. A.; Thörle-Pospiech, P.; Tinschert, K.; Trautmann, N.; Uusitalo, J.; Yeremin, A. V. (2016). "Remarks on the Fission Barriers of SHN and Search for Element 120". In Peninozhkevich, Yu. E.; Sobolev, Yu. G. Exotic Nuclei: EXON-2016 Proceedings of the International Symposium on Exotic Nuclei. Exotic Nuclei. pp. 155–164. ISBN 9789813226555.
  21. 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.
  22. "New Chemical Element In The Periodic Table". Science Daily. 11 June 2009.
  23. "Element 112 shall be named "copernicium"". Gesellschaft für Schwerionenforschung. 14 July 2009. Archived from the original on 18 July 2009.
  24. 1 2 "New element named 'copernicium'". BBC News. 16 July 2009. Retrieved 2010-02-22.
  25. "Start of the Name Approval Process for the Element of Atomic Number 112". IUPAC. July 20, 2009. Archived from the original on November 27, 2012. Retrieved April 14, 2012.
  26. Meija, J. (2009). "The need for a fresh symbol to designate copernicium". Nature. 461 (7262): 341. Bibcode:2009Natur.461..341M. doi:10.1038/461341c. PMID 19759598.
  27. van der Krogt, P. "Lutetium". Elementymology & Elements Multidict. Retrieved 2010-02-22.
  28. "IUPAC Element 112 is Named Copernicium". IUPAC. February 19, 2010. Archived from the original on March 4, 2016. Retrieved April 13, 2012.
  29. 1 2 3 4 5 Holden, N. E. (2004). "Table of the Isotopes". In D. R. Lide. CRC Handbook of Chemistry and Physics (85th ed.). CRC Press. Section 11. ISBN 978-0-8493-0485-9.
  30. 1 2 Dullman, C.E. Superheavy Element Research Superheavy Element - News from GSI and Mainz. University Mainz
  31. Barber, R. C.; et al. (2011). "Discovery of the elements with atomic numbers greater than or equal to 113" (PDF). Pure and Applied Chemistry. 83 (7): 5–7. doi:10.1351/PAC-REP-10-05-01.
  32. 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: Conference Series. 420. IOP Science. pp. 1–15. Retrieved 20 August 2013.
  33. Ninov, V.; et al. (1999). "86
    Kr
    ". Physical Review Letters. 83 (6): 1104–1107. Bibcode:1999PhRvL..83.1104N. doi:10.1103/PhysRevLett.83.1104.
  34. Public Affairs Department (July 21, 2001). "Results of element 118 experiment retracted". Berkeley Lab. Archived from the original on January 29, 2008. Retrieved January 18, 2008.
  35. At Lawrence Berkeley, Physicists Say a Colleague Took Them for a Ride George Johnson, The New York Times, 15 October 2002
  36. Public Affairs Department (26 October 2010). "Six New Isotopes of the Superheavy Elements Discovered: Moving Closer to Understanding the Island of Stability". Berkeley Lab. Retrieved 2011-04-25.
  37. 1 2 "Chemistry on the islands of stability", New Scientist, 11 September 1975, p. 574, ISSN 1032-1233
  38. Brändas, Erkki J.; Kryachko, Eugene S. (2013). Fundamental World of Quantum Chemistry. 3. Springer Science & Business Media. p. 348. ISBN 9789401704489.
  39. Demissie, Taye B.; Ruud, Kenneth (25 February 2017). "Darmstadtium, roentgenium, and copernicium form strong bonds with cyanide". International Journal of Quantum Chemistry. 2017: e25393. doi:10.1002/qua.25393.
  40. Kratz, Jens Volker. The Impact of Superheavy Elements on the Chemical and Physical Sciences. 4th International Conference on the Chemistry and Physics of the Transactinide Elements, 5 – 11 September 2011, Sochi, Russia
  41. 1 2 Gaston, Nicola; Opahle, Ingo; Gäggeler, Heinz W.; Schwerdtfeger, Peter (2007). "Is eka-mercury (element 112) a group 12 metal?". Angewandte Chemie. 46 (10): 1663–6. doi:10.1002/anie.200604262. Retrieved 5 November 2013.
  42. Čenčariková, Hana; Legut, Dominik (2017). "The effect of relativity on stability of Copernicium phases, their electronic structure and mechanical properties". Physica B. 536: 576. Bibcode:2018PhyB..536..576C. doi:10.1016/j.physb.2017.11.035.
  43. Borschevsky, Anastasia; Pershina, Valeria; Kaldor, Uzi; Eliav, Ephraim. "Fully relativistic ab initio studies of superheavy elements" (PDF). www.kernchemie.uni-mainz.de. Johannes Gutenberg University Mainz. Archived from the original (PDF) on January 15, 2018. Retrieved January 15, 2018.
  44. 1 2 3 4 5 Gäggeler, H. W. (2007). "Gas Phase Chemistry of Superheavy Elements" (PDF). Paul Scherrer Institute. pp. 26–28. Archived from the original (PDF) on 2012-02-20.
  45. Moody, Ken. "Synthesis of Superheavy Elements". In Schädel, Matthias; Shaughnessy, Dawn. The Chemistry of Superheavy Elements (2nd ed.). Springer Science & Business Media. pp. 24–8. ISBN 9783642374661.
  46. Zaitsevskii, A.; van Wüllen, C.; Rusakov, A.; Titov, A. (September 2007). "Relativistic DFT and ab initio calculations on the seventh-row superheavy elements: E113 - E114" (PDF). jinr.ru. Retrieved 17 February 2018.
  47. Paul Scherrer Institute (2015). "Annual Report 2015: Laboratory of Radiochemistry and Environmental Chemistry" (PDF). Paul Scherrer Institute. p. 3.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.