Jump to content

Meitnerium

This is a good article. Click here for more information.
From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Bibcode Bot (talk | contribs) at 04:54, 24 August 2013 (Adding 0 arxiv eprint(s), 2 bibcode(s) and 0 doi(s). Did it miss something? Report bugs, errors, and suggestions at User talk:Bibcode Bot). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Meitnerium, 109Mt
Meitnerium
Pronunciation
Mass number[278] (data not decisive)[a]
Meitnerium 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
Ir

Mt

hassiummeitneriumdarmstadtium
Atomic number (Z)109
Groupgroup 9
Periodperiod 7
Block  d-block
Electron configuration[Rn] 5f14 6d7 7s2 (predicted)[6][7]
Electrons per shell2, 8, 18, 32, 32, 15, 2 (predicted)
Physical properties
Phase at STPsolid (predicted)[8]
Density (near r.t.)27–28 g/cm3 (predicted)[9][10]
Atomic properties
Oxidation states(+1), (+3), (+4), (+6), (+8), (+9) (predicted)[6][11][12][13]
Ionization energies
  • 1st: 800 kJ/mol
  • 2nd: 1820 kJ/mol
  • 3rd: 2900 kJ/mol
  • (more) (all estimated)[6]
Atomic radiusempirical: 128 pm (predicted)[6][13]
Covalent radius129 pm (estimated)[14]
Other properties
Natural occurrencesynthetic
Crystal structureface-centered cubic (fcc)
Face-centered cubic crystal structure for meitnerium

(predicted)[8]
Magnetic orderingparamagnetic (predicted)[15]
CAS Number54038-01-6
History
Namingafter Lise Meitner
DiscoveryGesellschaft für Schwerionenforschung (1982)
Isotopes of meitnerium
Main isotopes[3] Decay
abun­dance half-life (t1/2) mode pro­duct
274Mt synth 0.64 s α 270Bh
276Mt synth 0.62 s α 272Bh
278Mt synth 4 s α 274Bh
282Mt synth 67 s?[5] α 278Bh
 Category: Meitnerium
| references

Meitnerium is a chemical element with symbol Mt and atomic number 109. It is an extremely radioactive synthetic element (an element not found in nature that can be created in a laboratory). The most stable known isotope, meitnerium-278, has a half-life of 7.6 seconds. The GSI Helmholtz Centre for Heavy Ion Research near Darmstadt, Germany, first created this element in 1982.

In the periodic table, is a d-block transactinide element. It is a member of the 7th period and is placed in the group 9 elements, although no chemical experiments have been carried out to confirm that it behaves as the heavier homologue to iridium in group 9. Meitnerium is calculated to have similar properties to its lighter homologues, cobalt, rhodium, and iridium.

History

Meitnerium was named after the physicist Lise Meitner, one of the discoverers of nuclear fission.

Discovery

Meitnerium was first synthesized on August 29, 1982 by a German research team led by Peter Armbruster and Gottfried Münzenberg at the Institute for Heavy Ion Research (Gesellschaft für Schwerionenforschung) in Darmstadt.[16] The team bombarded a target of bismuth-209 with accelerated nuclei of iron-58 and detected a single atom of the isotope meitnerium-266:

209
83
Bi
+ 58
26
Fe
266
109
Mt
+
n

Naming

The naming of meitnerium was discussed in the element naming controversy regarding the names of elements 104 to 109, but meitnerium was the only proposal and thus was never disputed.[17][18] The name meitnerium (Mt) was suggested in honor of the Austrian physicist Lise Meitner, a co-discoverer of protactinium (with Otto Hahn),[19][20][21][22][23] and one of the discoverers of nuclear fission.[24] In 1994 the name was recommended by IUPAC,[17] and was officially adopted in 1997.[18]

Nucleosynthesis

Super-heavy elements such as meitnerium are produced by bombarding lighter elements in particle accelerators that induce fusion reactions. Whereas the lightest isotope of meitnerium, meitnerium-266, can be synthesized directly this way, all the heavier meitnerium isotopes have only been observed as decay products of elements with higher atomic numbers.[25]

Depending on the energies involved, the former are separated into "hot" and "cold". In hot fusion reactions, very light, high-energy projectiles are accelerated toward very heavy targets (actinides), giving rise to compound nuclei at high excitation energy (~40–50 MeV) that may either fission or evaporate several (3 to 5) neutrons.[26] In cold fusion reactions, the produced fused nuclei have a relatively low excitation energy (~10–20 MeV), which decreases the probability that these products will undergo fission reactions. As the fused nuclei cool to the ground state, they require emission of only one or two neutrons, and thus, allows for the generation of more neutron-rich products.[25] The latter is a distinct concept from that of where nuclear fusion claimed to be achieved at room temperature conditions (see cold fusion).[27]

Cold fusion

After the first successful synthesis of meitnerium in 1982 by the GSI team,[16] a team at the Joint Institute for Nuclear Research in Dubna, Russia, also tried to observe the new element by bombarding bismuth-209 with iron-58. In 1985 they managed to identity alpha decays from the descendant isotope 246Cf indicating the formation of meitnerium. The observation of a further two atoms of 266Mt from the same reaction was reported in 1988 and of another 12 in 1997 by the German team at GSI.[28][29]

The same meitnerium isotope was also observed by the Russian team at Dubna in 1985 from the reaction:

208
82
Pb
+ 59
27
Co
266
109
Mt
+
n

by detecting the alpha decay of the descendant 246Cf nuclei. In 2007, an American team at the Lawrence Berkeley National Laboratory (LBNL) confirmed the decay chain of the 266Mt isotope from this reaction.[30]

Hot fusion

In 2002–2003, the team at the LBNL attempted to generate the isotope 271Mt to study its chemical properties by bombarding uranium-238 with chlorine-37, but without success.[31] Another possible reaction that would form this isotope would be the fusion of berkelium-249 with magnesium-26; however, the yield for this reaction is expected to be very low due to the high radioactivity of the berkelium-249 target.[6] Other long-lived isotopes were unsuccessfully targeted by a team at Lawrence Livermore National Laboratory (LLNL) in 1988 by bombarding einsteinium-254 with neon-22.[31]

Decay products

List of meitnerium isotopes observed by decay
Evaporation residue Observed meitnerium isotope
294Uus, 290Uup, 286Uut, 282Rg 278Mt[32]
288Uup, 284Uut, 280Rg 276Mt[33]
287Uup, 283Uut, 279Rg 275Mt[33]
282Uut, 278Rg 274Mt[33]
278Uut, 274Rg 270Mt[34]
272Rg 268Mt[35]

All the isotopes of meitnerium except meitnerium-266 have been detected only in the decay chains of elements with a higher atomic number, such as roentgenium. Roentgenium currently has seven known isotopes; all but one of them undergo alpha decays to become meitnerium nuclei, with mass numbers between 268 and 278. Parent roentgenium nuclei can be themselves decay products of ununtrium, ununpentium, or ununseptium. To date, no other elements have been known to decay to meitnerium.[36] For example, in January 2010, the Dubna team (JINR) identified meitnerium-278 as a product in the decay of ununseptium via an alpha decay sequence:[32]

The element ununseptium does not exist.The element ununpentium does not exist. + 4
2
He
The element ununpentium does not exist.The element ununtrium does not exist. + 4
2
He
The element ununtrium does not exist.282
111
Rg
+ 4
2
He
282
111
Rg
278
109
Mt
+ 4
2
He

Isotopes

List of meitnerium isotopes
Isotope
Half-life
[36][37]
Decay
mode[36][37]
Discovery
year
Reaction
265Mt 2? min α ? unknown
266Mt 1.7 ms α 1982 209Bi(58Fe,n)[16]
267Mt 10? ms α ? unknown
268Mt 21 ms α 1994 272Rg(—,α)[35]
269Mt 0.2? s α ? unknown
270Mt 5.0 ms α 2004 278Uut(—,2α)[33][34]
271Mt 5? s α ? unknown
272Mt 10? s α, SF ? unknown
273Mt 20? s α, SF ? unknown
274Mt 0.44 s α, SF 2006 282Uut(—,2α)[33]
275Mt 9.7 ms α 2003 287Uup(—,3α)[33]
276Mt 0.72 s α, SF 2003 288Uup(—,3α)[33]
277Mt 1? min α, SF ? unknown
278Mt 7.6 s α 2009 294Uus(—,4α)[32]
279Mt 6? min α, SF ? unknown

Meitnerium 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 of meitnerium have been reported with atomic masses 266, 268, 270, 274–276, and 278, two of which, meitnerium-268 and meitnerium-270, have known but unconfirmed metastable states. Most of these decay predominantly through alpha decay, although some undergo spontaneous fission.[36]

Stability and half-lives

All meitnerium isotopes are extremely unstable and radioactive; in general, heavier isotopes are more stable than the lighter. The most stable known meitnerium isotope, 278Mt, is also the heaviest known meitnerium isotope; it has a half-life of 7.6 seconds. A metastable nuclear isomer, 270mMt, has been reported to also have a half-life of over a second. The isotopes 276Mt and 274Mt have half-lives of 0.72 and 0.44 seconds respectively. The remaining four isotopes have half-lives between 1 and 20 milliseconds.[36] The undiscovered isotope 281Mt has been predicted to be the most stable towards beta decay;[38] however, no known meitnerium isotope has been observed to undergo beta decay.[36] Some unknown isotopes, such as 265Mt, 272Mt, 273Mt, 277Mt, and 279Mt, are predicted to have half-lives longer than the known isotopes.[36][37] Before its discovery, 274Mt was also predicted to have a long half-life of 20 seconds; however, it was later found to have a shorter half-life of only 0.44 seconds.[36]

Nuclear isomerism

270Mt

Two atoms of 270Mt have been identified in the decay chains of 278Uut. The two decays have very different lifetimes and decay energies and are also produced from two apparently different isomers of 274Rg. The first isomer decays by emission of an alpha particle with energy 10.03 MeV and has a lifetime of 7.16 ms. The other alpha decays with a lifetime of 1.63 s; the decay energy was not measured. An assignment to specific levels is not possible with the limited data available and further research is required.[34]

268Mt

The alpha decay spectrum for 268Mt appears to be complicated from the results of several experiments. Alpha particles of energies 10.28, 10.22 and 10.10 MeV have been observed, emitted from 268Mt atoms with half-lives of 42 ms, 21 ms and 102 ms respectively. The long-lived decay must be assigned to an isomeric level. The discrepancy between the other two half-lives has yet to be resolved. An assignment to specific levels is not possible with the data available and further research is required.[35]

Predicted properties

Chemical

Meitnerium is the seventh member of the 6d series of transition metals. Since element 112 (copernicium) has been shown to be a transition metal, it is expected that all the elements from 104 to 112 would form a fourth transition metal series, with meitnerium as part of the platinum group metals.[22] Calculations on its ionization potentials and atomic and ionic radii are similar to that of its lighter homologue iridium, thus implying that meitnerium's basic properties will resemble those of the other group 9 elements, cobalt, rhodium, and iridium.[6]

Prediction of the probable chemical properties of meitnerium has not received much attention recently. Meitnerium is expected to be a noble metal. Based on the most stable oxidation states of the lighter group 9 elements, the most stable oxidation states of meitnerium are predicted to be the +6, +3, and +1 states, with the +3 state being the most stable in aqueous solutions. In comparison, rhodium and iridium show a maximum oxidation state of +6, while the most stable states are +4 and +3 for iridium and +3 for rhodium.[6] Group 9 is the first group in the transition metals to show lower oxidation states than the group number, the +9 state not being known for any element. The oxidation state +9 might be possible for meitnerium in the nonafluoride (MtF9) and the [MtO4]+ cation, although [IrO4]+ is expected to be more stable.[12] The tetrahalides of meitnerium have also been predicted to have similar stabilities to those of iridium, thus also allowing a stable +4 state.[11] It is further expected that the maximum oxidation states of elements from bohrium (element 107) to darmstadtium (element 110) may be stable in the gas phase but not in aqueous solution.[6]

Physical and atomic

Meitnerium is expected to be a solid under normal conditions and assume a face-centered cubic crystal structure, similarly to its lighter congener iridium.[8] It should be a very heavy metal with a density of around 37.4 g/cm3, which would be the second-highest of any of the 118 known elements, second only to that predicted for its neighbor hassium (41 g/cm3). In comparison, the densest known element that has had its density measured, osmium, has a density of only 22.61 g/cm3. This results from meitnerium's high atomic weight, the lanthanide and actinide contractions, and relativistic effects, although production of enough meitnerium to measure this quantity would be impractical, and the sample would quickly decay.[6] Meitnerium is also predicted to be paramagnetic.[15]

Theoreticians have predicted the covalent radius of meitnerium to be 6 to 10 pm larger than that of iridium.[39] For example, the Mt–O bond distance is expected to be around 1.9 Å.[40] The atomic radius of meitnerium is expected to be around 122 pm.[6]

Experimental chemistry

Unambiguous determination of the chemical characteristics of meitnerium has yet to have been established[41][42] due to the short half-lives of meitnerium isotopes[6] and a limited number of likely volatile compounds that could be studied on a very small scale. One of the few meitnerium compounds that are likely to be sufficiently volatile is meitnerium hexafluoride (MtF
6
), as its lighter homologue iridium hexafluoride (IrF
6
) is volatile above 60 °C and therefore the analogous compound of meitnerium might also be sufficiently volatile;[22] a volatile octafluoride (MtF
8
) might also be possible.[6] For chemical studies to be carried out on a transactinide, at least four atoms must be produced, the half-life of the isotope used must be at least 1 second, and the rate of production must be at least one atom per week.[22] Even though the half-life of 278Mt, the most stable known meitnerium isotope, is 7.6 seconds, long enough to perform chemical studies, another obstacle is the need to increase the rate of production of meitnerium isotopes and allow experiments to carry on for weeks or months so that statistically significant results can be obtained. Separation and detection must be carried out continuously to separate out the meitnerium isotopes and automated systems can then experiment on the gas-phase and solution chemistry of meitnerium as the yields for heavier elements are predicted to be smaller than those for lighter elements; some of the separation techniques used for bohrium and hassium could be reused. However, the experimental chemistry of meitnerium has not received as much attention as that of the heavier elements copernicium and flerovium.[6][41]

The Lawrence Berkeley National Laboratory attempted to synthesize the isotope 271Mt in 2002–2003 for a possible chemical investigation of meitnerium because it was expected that it might be more stable than the isotopes around it as it has 162 neutrons, a magic number for deformed nuclei; its half-life was predicted to be a few seconds, long enough for a chemical investigation.[6][43] However, no atoms of 271Mt were detected,[31] and this isotope of meitnerium is currently unknown.[36]

An experiment determining the chemical properties of a transactinide would need to compare a compound of that transactinide with analogous compounds of some of its lighter homologues:[6] for example, in the chemical characterization of hassium, hassium tetroxide (HsO4) was compared with the analogous osmium compound, osmium tetroxide (OsO4).[44] In a preliminary step towards determining the chemical properties of meitnerium, the GSI attempted sublimation of the rhodium compounds rhodium(III) oxide (Rh2O3) and rhodium(III) chloride (RhCl3). However, macroscopic amounts of the oxide would not sublimate until 1000 °C and the chloride would not until 780 °C, and then only in the presence of carbon aerosol particles: these temperatures are far too high for such procedures to be used on meitnerium, as most of the current methods used for the investigation of the chemistry of superheavy elements do not work above 500 °C.[42]

References

  1. ^ Emsley, John (2003). Nature's Building Blocks. Oxford University Press. ISBN 978-0198503408. Retrieved November 12, 2012.
  2. ^ Meitnerium. The Periodic Table of Videos. University of Nottingham. February 18, 2010. Retrieved October 15, 2012.
  3. ^ a b Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  4. ^ Oganessian, Yu. Ts.; Utyonkov, V. K.; Kovrizhnykh, N. D.; et al. (2022). "New isotope 286Mc produced in the 243Am+48Ca reaction". Physical Review C. 106 (64306): 064306. Bibcode:2022PhRvC.106f4306O. doi:10.1103/PhysRevC.106.064306. S2CID 254435744.
  5. ^ a b 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.; Scheidenberger, C.; Schött, H. J.; Shaughnessy, D. A.; Stoyer, M. A.; Thörle-Popiesch, P.; Tinschert, K.; Trautmann, N.; Uusitalo, J.; Yeremin, A. V. (2016). "Review of even element super-heavy nuclei and search for element 120". The European Physics Journal A. 2016 (52). doi:10.1140/epja/i2016-16180-4.
  6. ^ a b c d e f g h i j k l m n o 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.
  7. ^ Thierfelder, C.; Schwerdtfeger, P.; Heßberger, F. P.; Hofmann, S. (2008). "Dirac-Hartree-Fock studies of X-ray transitions in meitnerium". The European Physical Journal A. 36 (2): 227. Bibcode:2008EPJA...36..227T. doi:10.1140/epja/i2008-10584-7.
  8. ^ a b c Östlin, A.; Vitos, L. (2011). "First-principles calculation of the structural stability of 6d transition metals". Physical Review B. 84 (11): 113104. Bibcode:2011PhRvB..84k3104O. doi:10.1103/PhysRevB.84.113104.
  9. ^ Gyanchandani, Jyoti; Sikka, S. K. (10 May 2011). "Physical properties of the 6 d -series elements from density functional theory: Close similarity to lighter transition metals". Physical Review B. 83 (17): 172101. Bibcode:2011PhRvB..83q2101G. doi:10.1103/PhysRevB.83.172101.
  10. ^ Kratz; Lieser (2013). Nuclear and Radiochemistry: Fundamentals and Applications (3rd ed.). p. 631.
  11. ^ a b Ionova, G. V.; Ionova, I. S.; Mikhalko, V. K.; Gerasimova, G. A.; Kostrubov, Yu. N.; Suraeva, N. I. (2004). "Halides of Tetravalent Transactinides (Rf, Db, Sg, Bh, Hs, Mt, 110th Element): Physicochemical Properties". Russian Journal of Coordination Chemistry. 30 (5): 352. doi:10.1023/B:RUCO.0000026006.39497.82. S2CID 96127012.
  12. ^ a b Himmel, Daniel; Knapp, Carsten; Patzschke, Michael; Riedel, Sebastian (2010). "How Far Can We Go? Quantum-Chemical Investigations of Oxidation State +IX". ChemPhysChem. 11 (4): 865–9. doi:10.1002/cphc.200900910. PMID 20127784.
  13. ^ a b Fricke, Burkhard (1975). "Superheavy elements: a prediction of their chemical and physical properties". Recent Impact of Physics on Inorganic Chemistry. Structure and Bonding. 21: 89–144. doi:10.1007/BFb0116498. ISBN 978-3-540-07109-9. Retrieved 4 October 2013.
  14. ^ Chemical Data. Meitnerium - Mt, Royal Chemical Society
  15. ^ a b Saito, Shiro L. (2009). "Hartree–Fock–Roothaan energies and expectation values for the neutral atoms He to Uuo: The B-spline expansion method". Atomic Data and Nuclear Data Tables. 95 (6): 836–870. Bibcode:2009ADNDT..95..836S. doi:10.1016/j.adt.2009.06.001.
  16. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1007/BF01420157, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1007/BF01420157 instead.
  17. ^ a b "Names and symbols of transfermium elements (IUPAC Recommendations 1994)". Pure and Applied Chemistry. 66 (12): 2419. 1994. doi:10.1351/pac199466122419.
  18. ^ a b Rayner-Canham, Geoff; Zheng, Zheng (2007). "Naming elements after scientists: An account of a controversy". Foundations of Chemistry. 10: 13. doi:10.1007/s10698-007-9042-1.
  19. ^ Attention: This template ({{cite pmid}}) is deprecated. To cite the publication identified by PMID 11206992, please use {{cite journal}} with |pmid=11206992 instead.
  20. ^ Attention: This template ({{cite pmid}}) is deprecated. To cite the publication identified by PMID 7014939, please use {{cite journal}} with |pmid=7014939 instead.
  21. ^ Attention: This template ({{cite pmid}}) is deprecated. To cite the publication identified by PMID 4573793, please use {{cite journal}} with |pmid=4573793 instead.
  22. ^ a b c d Griffith, W. P. (2008). "The Periodic Table and the Platinum Group Metals". Platinum Metals Review. 52 (2): 114. doi:10.1595/147106708X297486.
  23. ^ Rife, Patricia (2003). "Meitnerium". Chemical & Engineering News. 81 (36): 186. doi:10.1021/cen-v081n036.p186.
  24. ^ Wiesner, Emilie; Settle, Frank A. (2001). "Politics, Chemistry, and the Discovery of Nuclear Fission". Journal of Chemical Education. 78 (7): 889. Bibcode:2001JChEd..78..889W. doi:10.1021/ed078p889.
  25. ^ a b Armbruster, Peter; Munzenberg, Gottfried (1989). "Creating superheavy elements". Scientific American. 34: 36–42. {{cite journal}}: Unknown parameter |lastauthoramp= ignored (|name-list-style= suggested) (help)
  26. ^ Barber, Robert C.; Gäggeler, Heinz W.; Karol, Paul J.; Nakahara, Hiromichi; Vardaci, Emanuele; Vogt, Erich (2009). "Discovery of the element with atomic number 112 (IUPAC Technical Report)". Pure and Applied Chemistry. 81 (7): 1331. doi:10.1351/PAC-REP-08-03-05.
  27. ^ Fleischmann, Martin; Pons, Stanley (1989). "Electrochemically induced nuclear fusion of deuterium". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 261 (2): 301–308. doi:10.1016/0022-0728(89)80006-3.
  28. ^ Münzenberg, G.; Hofmann, S.; Heßberger, F. P.; et al. (1988). "New results on element 109". Zeitschrift für Physik A: Atoms and Nuclei. 330 (4): 435–436. Bibcode:1988ZPhyA.330..435M. doi:10.1007/BF01290131. {{cite journal}}: replacement character in |last9= at position 4 (help)
  29. ^ Hofmann, S.; Heßberger, F. P.; Ninov, V.; et al. (1997). "Excitation function for the production of 265108 and 266109". Zeitschrift für Physik A: Atoms and Nuclei. 358 (4): 377–378. Bibcode:1997ZPhyA.358..377H. doi:10.1007/s002180050343.
  30. ^ Nelson, S. L.; Gregorich, K. E.; Dragojević, I.; et al. (2009). "Comparison of complementary reactions in the production of Mt". Physical Review C. 79 (2): 027605. Bibcode:2009PhRvC..79b7605N. doi:10.1103/PhysRevC.79.027605.
  31. ^ a b c Zielinski P. M. et al. (2003). "The search for 271Mt via the reaction 238U + 37Cl", GSI Annual report. Retrieved on 2008-03-01
  32. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1103/PhysRevLett.104.142502, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1103/PhysRevLett.104.142502 instead.
  33. ^ a b c d e f g Oganessian, Yu. Ts.; Penionzhkevich, Yu. E.; Cherepanov, E. A. (2007). "AIP Conference Proceedings". 912: 235. doi:10.1063/1.2746600. {{cite journal}}: |chapter= ignored (help); Cite journal requires |journal= (help)
  34. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1143/JPSJ.73.2593, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1143/JPSJ.73.2593 instead.
  35. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1007/BF01291182, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1007/BF01291182 instead.
  36. ^ a b c d e f g h i Sonzogni, Alejandro. "Interactive Chart of Nuclides". National Nuclear Data Center: Brookhaven National Laboratory. Retrieved 2008-06-06.
  37. ^ a b c Gray, Theodore (2002–2010). "The Photographic Periodic Table of the Elements". periodictable.com. Retrieved 16 November 2012.
  38. ^ Nie, G. K. (2005). "Charge radii of β-stable nuclei". Modern Physics Letters A. 21 (24): 1889. arXiv:nucl-th/0512023. Bibcode:2006MPLA...21.1889N. doi:10.1142/S0217732306020226.
  39. ^ Pyykkö, Pekka; Atsumi, Michiko (2009). "Molecular Double-Bond Covalent Radii for Elements Li—E112". Chemistry – A European Journal. 15 (46): 12770. doi:10.1002/chem.200901472.
  40. ^ Van Lenthe, E.; Baerends, E. J. (2003). "Optimized Slater-type basis sets for the elements 1–118". Journal of Computational Chemistry. 24 (9): 1142–56. doi:10.1002/jcc.10255. PMID 12759913.
  41. ^ a b Düllmann, Christoph E. (2012). "Superheavy elements at GSI: a broad research program with element 114 in the focus of physics and chemistry". Radiochimica Acta. 100 (2): 67–74. doi:10.1524/ract.2011.1842.
  42. ^ a b Haenssler, F. L.; Düllmann, Ch. E.; Gäggeler, H. W.; Eichler, B. "Thermatographic investigation of Rh and 107Rh with different carrier gases" (PDF). Retrieved 15 October 2012.{{cite web}}: CS1 maint: multiple names: authors list (link)
  43. ^ Smolańczuk, R. (1997). "Properties of the hypothetical spherical superheavy nuclei". Phys. Rev. C. 56 (2): 812–24. Bibcode:1997PhRvC..56..812S. doi:10.1103/PhysRevC.56.812.
  44. ^ Düllmann, Ch. E for a Univ. Bern - PSI - GSI - JINR - LBNL - Univ. Mainz - FZR - IMP - collaboration. "Chemical investigation of hassium (Hs, Z=108)" (PDF). Retrieved 15 October 2012.

Template:Chemical elements named after scientists
Cite error: There are <ref group=lower-alpha> tags or {{efn}} templates on this page, but the references will not show without a {{reflist|group=lower-alpha}} template or {{notelist}} template (see the help page).