Moscovium

"Element 115" redirects here. For fictional and conspiracy references to element 115, see Materials science in science fiction.

Moscovium,  115Mc
General properties
Name, symbol moscovium, Mc
Pronunciation /mɒsˈkviəm/
mos-KOH-vee-əm
Moscovium in the periodic table
Bi

Mc

(Uhe)
fleroviummoscoviumlivermorium
Atomic number (Z) 115
Group, block group 15 (pnictogens), p-block
Period period 7
Element category unknown, but probably a post-transition metal
Standard atomic weight (Ar) [289]
Electron configuration [Rn] 5f14 6d10 7s2 7p3 (predicted)[1]
per shell
2, 8, 18, 32, 32, 18, 5 (predicted)
Physical properties
Phase solid (predicted)[1]
Melting point 670 K (400 °C, 750 °F) (predicted)[1][2]
Boiling point ~1400 K (~1100 °C, ~2000 °F) (predicted)[1]
Density near r.t. 13.5 g/cm3 (predicted)[2]
Heat of fusion 5.90–5.98 kJ/mol (extrapolated)[3]
Heat of vaporization 138 kJ/mol (predicted)[2]
Atomic properties
Oxidation states 1, 3 (prediction)[1][2]
Ionization energies 1st: 538.4 kJ/mol (predicted)[1]
2nd: 1756.0 kJ/mol (predicted)[2]
3rd: 2653.3 kJ/mol (predicted)[2]
(more)
Atomic radius empirical: 187 pm (predicted)[1][2]
Covalent radius 156–158 pm (extrapolated)[3]
Miscellanea
CAS Number 54085-64-2
History
Naming After Moscow region
Discovery Joint Institute for Nuclear Research and Lawrence Livermore National Laboratory (2003)
Most stable isotopes of moscovium
iso NA half-life DM DE (MeV) DP
290Mc syn 16 ms[4] α 9.95 286Nh
289Mc syn 220 ms[4] α 10.31 285Nh
288Mc syn 87 ms α 10.46 284Nh
287Mc syn 32 ms α 10.59 283Nh

Moscovium (element symbol: Mc) is a superheavy synthetic element that was first synthesized in 2003 by a joint team of Russian and American scientists at the Joint Institute for Nuclear Research (JINR) in Dubna, Russia. With the atomic number 115 in the periodic table, it is also known as element 115, or eka-bismuth. In December 2015, it was recognized as one of four new elements by the Joint Working Party of international scientific bodies IUPAC and IUPAP. On 28 November 2016, it was officially named.[5][6][7]

Moscovium is an extremely radioactive element: its most stable known isotope, moscovium-289, has a half-life of only 220 milliseconds.[8] In the periodic table, it is a p-block transactinide element. It is a member of the 7th period and is placed in group 15 as the heaviest pnictogen, although it has not been confirmed to behave as a heavier homologue of the pnictogen bismuth. Moscovium is calculated to have some properties similar to its lighter homologues, nitrogen, phosphorus, arsenic, antimony, and bismuth, and to be a post-transition metal, although it should also show several major differences from them. About 100 atoms of moscovium have been observed to date, all of which have been shown to have mass numbers from 287 to 290.

History

Discovery

The first successful synthesis of moscovium was by a joint team of Russian and American scientists in August 2003 at the Joint Institute for Nuclear Research (JINR) in Dubna, Russia. Headed by Russian nuclear physicist Yuri Oganessian, the team included American scientists of the Lawrence Livermore National Laboratory. The researchers on February 2, 2004, stated in Physical Review C that they bombarded americium-243 with calcium-48 ions to produce four atoms of moscovium. These atoms decayed by emission of alpha-particles to nihonium in about 100 milliseconds.[9][10]

243
95
Am
+ 48
20
Ca
288
115
Mc
+ 3 1
0
n
284
113
Nh
+
α
243
95
Am
+ 48
20
Ca
287
115
Mc
+ 4 1
0
n
283
113
Nh
+
α

The Dubna–Livermore collaboration strengthened their claim to the discoveries of moscovium and nihonium by conducting chemical experiments on the final decay product 268Db. None of the nuclides in this decay chain were previously known, so existing experimental data was not available to support their claim. In June 2004 and December 2005, the presence of a dubnium isotope was confirmed by extracting the final decay products, measuring spontaneous fission (SF) activities and using chemical identification techniques to confirm that they behave like a group 5 element (as dubnium is known to be in group 5 of the periodic table).[1][11] Both the half-life and the decay mode were confirmed for the proposed 268Db, lending support to the assignment of the parent nucleus to moscovium.[11][12] However, in 2011, the IUPAC/IUPAP Joint Working Party (JWP) did not recognize the two elements as having been discovered, because current theory could not distinguish the chemical properties of group 4 and group 5 elements with sufficient confidence.[13] Furthermore, the decay properties of all the nuclei in the decay chain of moscovium had not been previously characterized before the Dubna experiments, a situation which the JWP generally considers "troublesome, but not necessarily exclusive".[13]

Road to confirmation

Two heavier isotopes of moscovium, 289Mc and 290Mc, were discovered in 2009–2010 as daughters of the tennessine isotopes 293Ts and 294Ts.[8] As of 2016 the JINR also has future plans to study lighter isotopes of moscovium by replacing the americium-243 target with the lighter isotope americium-241.[14]

In 2011, the Joint Working Party of international scientific bodies International Union of Pure and Applied Chemistry (IUPAC) and International Union of Pure and Applied Physics (IUPAP) evaluated the 2004 and 2007 Dubna experiments, and concluded that they did not meet the criteria for discovery. Another evaluation of more recent experiments took place within the next few years, and a claim to the discovery of moscovium was again put forward by Dubna.[13] In August 2013, a team of researchers at Lund University and at the Gesellschaft für Schwerionenforschung (GSI) in Darmstadt, Germany announced they had repeated the 2004 experiment, confirming Dubna's findings.[15][16] Further confirmation was published by the team at Berkeley in 2015.[17]

In December 2015, the IUPAC/IUPAP Joint Working Party recognized the element's discovery and assigned the priority to the Dubna-Berkeley collaboration of 2009–2010, giving them the right to suggest a permanent name for it.[18]

Naming

Using Mendeleev's nomenclature for unnamed and undiscovered elements, moscovium is sometimes known as eka-bismuth. In 1979 IUPAC recommended that the placeholder systematic element name ununpentium (with the corresponding symbol of Uup)[19] be used until the discovery of the element is confirmed and a permanent name is decided. Although widely used in the chemical community on all levels, from chemistry classrooms to advanced textbooks, the recommendations are mostly ignored among scientists in the field, who call it "element 115", with the symbol of (115) or even simply 115.[1]

On 30 December 2015, discovery of the element was recognized by the International Union of Pure and Applied Chemistry (IUPAC).[20] According to IUPAC recommendations, the discoverer(s) of a new element has the right to suggest a name.[21] A suggested name was langevinium, after Paul Langevin.[22] Later, the Dubna team mentioned the name moscovium several times as one among many possibilities, referring to the Moscow Oblast where Dubna is located.[23][24]

Wikinews has related news: IUPAC proposes four new chemical element names

In June 2016, IUPAC endorsed the latter proposal to be formally accepted by the end of the year.[7]

Predicted properties

Nuclear stability and isotopes

The expected location of the island of stability. The dotted line is the line of beta stability.

Moscovium is expected to be in the middle of an island of stability centered on copernicium (element 112) and flerovium (element 114): the reasons for the presence of this island, however, are still not well understood.[25][26] Due to the expected high fission barriers, any nucleus within this island of stability exclusively decays by alpha decay and perhaps some electron capture and beta decay.[2] Although the known isotopes of moscovium do not actually have enough neutrons to be on the island of stability, they can be seen to approach the island as in general, the heavier isotopes are the longer-lived ones.[8][11]

The hypothetical isotope 291Mc is an especially interesting case as it has only one neutron more than the heaviest known moscovium isotope, 290Mc. It could plausibly be synthesized as the daughter of tennessine-295, which in turn could be made from the reaction 249Bk(48Ca,2n)295Ts.[25] Calculations show that it may have a significant electron capture or positron emission decay mode in addition to alpha decaying and also have a relatively long half-life of several seconds. This would produce 291Fl, 291Nh, and finally 291Cn which is expected to be in the middle of the island of stability and have a half-life of about 1200 years, affording the most likely hope of reaching the middle of the island using current technology. Possible drawbacks are that the cross section of the production reaction of 295Ts is expected to be low and the decay properties of superheavy nuclei this close to the line of beta stability are largely unexplored.[25]

Other possibilities to synthesize nuclei on the island of stability include quasifission (partial fusion followed by fission) of a massive nucleus.[27] Such nuclei tend to fission, expelling doubly magic or nearly doubly magic fragments such as calcium-40, tin-132, lead-208, or bismuth-209.[28] Recently it has been shown that the multi-nucleon transfer reactions in collisions of actinide nuclei (such as uranium and curium) might be used to synthesize the neutron-rich superheavy nuclei located at the island of stability,[27] although formation of the lighter elements nobelium or seaborgium is more favored.[25] One last possibility to synthesize isotopes near the island is to use controlled nuclear explosions to create a neutron flux high enough to bypass the gaps of instability at 258–260Fm and at mass number 275 (atomic numbers 104 to 108), mimicking the r-process in which the actinides were first produced in nature and the gap of instability around radon bypassed.[25] Some such isotopes (especially 291Cn and 293Cn) may even have been synthesized in nature, but would have decayed away far too quickly (with half-lives of only thousands of years) and be produced in far too small quantities (about 10−12 the abundance of lead) to be detectable as primordial nuclides today outside cosmic rays.[25]

Physical and atomic

In the periodic table, moscovium is a member of group 15, the pnictogens, below nitrogen, phosphorus, arsenic, antimony, and bismuth. Every previous pnictogen has five electrons in its valence shell, forming a valence electron configuration of ns2np3. In moscovium's case, the trend should be continued and the valence electron configuration is predicted to be 7s27p3;[1] therefore, moscovium will behave similarly to its lighter congeners in many respects. However, notable differences are likely to arise; a largely contributing effect is the spin–orbit (SO) interaction—the mutual interaction between the electrons' motion and spin. It is especially strong for the superheavy elements, because their electrons move much faster than in lighter atoms, at velocities comparable to the speed of light.[29] In relation to moscovium atoms, it lowers the 7s and the 7p electron energy levels (stabilizing the corresponding electrons), but two of the 7p electron energy levels are stabilized more than the other four.[30] The stabilization of the 7s electrons is called the inert pair effect, and the effect "tearing" the 7p subshell into the more stabilized and the less stabilized parts is called subshell splitting. Computation chemists see the split as a change of the second (azimuthal) quantum number l from 1 to 12 and 32 for the more stabilized and less stabilized parts of the 7p subshell, respectively.[29][lower-alpha 1] For many theoretical purposes, the valence electron configuration may be represented to reflect the 7p subshell split as 7s2
7p2
1/2
7p1
3/2
.[1] These effects cause moscovium's chemistry to be somewhat different from that of its lighter congeners.

The valence electrons of moscovium fall into three subshells: 7s (two electrons), 7p1/2 (two electrons), and 7p3/2 (one electron). The first two of these are relativistically stabilized and hence behave as inert pairs, while the last is relativistically destabilized and can easily participate in chemistry.[1] (The 6d electrons are not destabilized enough to participate chemically, although this may still be possible in the two previous elements nihonium and flerovium.)[2] Thus, the +1 oxidation state should be favored, like Tl+, and consistent with this the first ionization potential of moscovium should be around 5.58 eV, continuing the trend towards lower ionization potentials down the pnictogens.[1] Moscovium and nihonium both have one electron outside a quasi-closed shell configuration that can be delocalized in the metallic state: thus they should have similar melting and boiling points (both melting around 400 °C and boiling around 1100 °C) due to the strength of their metallic bonds being similar.[2] Additionally, the predicted ionization potential, ionic radius (1.5 Å for Mc+; 1.0 Å for Mc3+), and polarizability of Mc+ are expected to be more similar to Tl+ than its true congener Bi3+.[2] Moscovium should be a dense metal due to its high atomic weight, with a density around 13.5 g/cm3.[2] The electron of the hydrogen-like moscovium atom (oxidized so that it only has one electron, Mc114+) is expected to move so fast that it has a mass 1.82 times that of a stationary electron, due to relativistic effects. For comparison, the figures for hydrogen-like bismuth and antimony are expected to be 1.25 and 1.077 respectively.[29]

Chemical

Moscovium is predicted to be the third member of the 7p series of chemical elements and the heaviest member of group 15 (VA) in the Periodic Table, below bismuth. In this group, each member is known to portray the group oxidation state of +5 but with differing stability. For nitrogen, the +5 state is mostly a formal explanation of molecules like N2O5: it is very difficult to have five covalent bonds to nitrogen due to the inability of the small nitrogen atom to accommodate five ligands. The +5 state is well represented for the essentially non-relativistic typical pnictogens phosphorus, arsenic, and antimony. However, for bismuth it becomes rare due to the relativistic stabilization of the 6s orbitals known as the inert pair effect, so that the 6s electrons are reluctant to bond chemically. It is expected that moscovium will have an inert pair effect for both the 7s and the 7p1/2 electrons, as the binding energy of the lone 7p3/2 electron is noticeably lower than that of the 7p1/2 electrons. Nitrogen(I) and bismuth(I) are known but rare and moscovium(I) is likely to show some unique properties,[31] probably behaving more like thallium(I) than bismuth(I).[2] Because of spin-orbit coupling, flerovium may display closed-shell or noble gas-like properties; if this is the case, moscovium will likely be typically monovalent as a result, since the cation Mc+ will have the same electron configuration as flerovium, perhaps giving moscovium some alkali metal character.[2] However, the Mc3+ cation would behave like its true lighter homolog Bi3+.[2] The 7s electrons are too stabilized to be able to contribute chemically and hence the +5 state should be impossible and moscovium may be considered to have only three valence electrons.[2] Moscovium would be quite a reactive metal, with a standard reduction potential of −1.5 V for the Mc+/Mc couple.[2]

The chemistry of moscovium in aqueous solution should essentially be that of the Mc+ and Mc3+ ions. The former should be easily hydrolyzed and not be easily complexed with halides, cyanide, and ammonia.[2] Moscovium(I) hydroxide (McOH), carbonate (Mc2CO3), oxalate (Mc2C2O4), and fluoride (McF) should be soluble in water; the sulfide (Mc2S) should be insoluble; and the chloride (McCl), bromide (McBr), iodide (McI), and thiocyanate (McSCN) should be only slightly soluble, so that adding excess hydrochloric acid would not noticeably affect the solubility of moscovium(I) chloride.[2] Mc3+ should be about as stable as Tl3+ and hence should also be an important part of moscovium chemistry, although its closest homolog among the elements should be its lighter congener Bi3+.[2] Moscovium(III) fluoride (McF3) and thiozonide (McS3) should be insoluble in water, similar to the corresponding bismuth compounds, while moscovium(III) chloride (McCl3), bromide (McBr3), and iodide (McI3) should be readily soluble and easily hydrolyzed to form oxyhalides such as McOCl and McOBr, again analogous to bismuth.[2] Both moscovium(I) and moscovium(III) should be common oxidation states and their relative stability should depend greatly on what they are complexed with and the likelihood of hydrolysis.[2]

Experimental chemistry

Unambiguous determination of the chemical characteristics of moscovium has yet to have been established.[32][33] In 2011, experiments were conducted to create nihonium, flerovium, and moscovium isotopes in the reactions between calcium-48 projectiles and targets of americium-243 and plutonium-244. However, the targets included lead and bismuth impurities and hence some isotopes of bismuth and polonium were generated in nucleon transfer reactions. This, while an unforeseen complication, could give information that would help in the future chemical investigation of the heavier homologs of bismuth and polonium, which are respectively moscovium and livermorium.[33] The produced nuclides bismuth-213 and polonium-212m were transported as the hydrides 213BiH3 and 212mPoH2 at 850 °C through a quartz wool filter unit held with tantalum, showing that these hydrides were surprisingly thermally stable, although their heavier congeners McH3 and LvH2 would be expected to be less thermally stable from simple extrapolation of periodic trends in the p-block.[33] Further calculations on the stability and electronic structure of BiH3, McH3, PoH2, and LvH2 are needed before chemical investigations take place. However, moscovium and livermorium are expected to be volatile enough as pure elements for them to be chemically investigated in the near future. The chief barrier to their chemical investigation at present is the lack of known isotopes of these elements which are long-lived enough, with only 289Mc being barely usable with current methods.[33]

Notes

  1. The quantum number corresponds to the letter in the electron orbital name: 0 to s, 1 to p, 2 to d, etc. See azimuthal quantum number for more information.

See also

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 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. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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.
  3. 1 2 Bonchev, Danail; Kamenska, Verginia (1981). "Predicting the Properties of the 113–120 Transactinide Elements". Journal of Physical Chemistry. American Chemical Society. 85 (9): 1177–1186. doi:10.1021/j150609a021.
  4. 1 2 Oganessian, Yuri Ts.; Abdullin, F. Sh.; Bailey, P. D.; et al. (2010-04-09). "Synthesis of a New Element with Atomic Number Z=117". Physical Review Letters. American Physical Society. 104 (142502). Bibcode:2010PhRvL.104n2502O. doi:10.1103/PhysRevLett.104.142502. PMID 20481935.
  5. Staff (30 November 2016). "IUPAC Announces the Names of the Elements 113, 115, 117, and 118". IUPAC. Retrieved 1 December 2016.
  6. St. Fleur, Nicholas (1 December 2016). "Four New Names Officially Added to the Periodic Table of Elements". New York Times. Retrieved 1 December 2016.
  7. 1 2 "IUPAC Is Naming The Four New Elements Nihonium, Moscovium, Tennessine, And Oganesson". IUPAC. 2016-06-08. Retrieved 2016-06-08.
  8. 1 2 3 Oganessian, Yuri Ts.; Abdullin, F. Sh.; Bailey, P. D.; et al. (2010-04-09). "Synthesis of a New Element with Atomic Number Z=117". Physical Review Letters. American Physical Society. 104 (142502). Bibcode:2010PhRvL.104n2502O. doi:10.1103/PhysRevLett.104.142502. PMID 20481935.
  9. Oganessian, Yu. Ts.; Utyonkoy, V. K.; Lobanov, Yu. V.; et al. (2004). "Experiments on the synthesis of element 115 in the reaction 243Am(48Ca,xn)291−x115". Physical Review C. 69 (2): 021601. Bibcode:2004PhRvC..69b1601O. doi:10.1103/PhysRevC.69.021601.
  10. Oganessian; et al. (2003). "Experiments on the synthesis of element 115 in the reaction 243Am(48Ca,xn)291−x115" (PDF). JINR preprints.
  11. 1 2 3 "Results of the experiment on chemical identification of Db as a decay product of element 115", Oganessian et al., JINR preprints, 2004. Retrieved on 3 March 2008
  12. Oganessian, Yu. Ts.; Utyonkov, V.; Dmitriev, S.; Lobanov, Yu.; Itkis, M.; Polyakov, A.; Tsyganov, Yu.; Mezentsev, A.; Yeremin, A.; Voinov, A. A.; et al. (2005). "Synthesis of elements 115 and 113 in the reaction 243Am + 48Ca". Physical Review C. 72 (3): 034611. Bibcode:2005PhRvC..72c4611O. doi:10.1103/PhysRevC.72.034611.
  13. 1 2 3 Barber, Robert C.; Karol, Paul J; Nakahara, Hiromichi; Vardaci, Emanuele; Vogt, Erich W. (2011). "Discovery of the elements with atomic numbers greater than or equal to 113 (IUPAC Technical Report)". Pure Appl. Chem. 83 (7): 1485. doi:10.1351/PAC-REP-10-05-01.
  14. "Study of heavy and superheavy nuclei (see project 1.5)". Flerov Laboratory of Nuclear Reactions.
  15. "Existence of new element confirmed". Lund University. 27 August 2013. Retrieved 10 April 2016.
  16. "Spectroscopy of element 115 decay chains (Accepted for publication on Physical Review Letters on 9 August 2013)". Retrieved 2 September 2013.
  17. "Decay spectroscopy of element 115 daughters: 280Rg→276Mt and 276Mt→Bh". doi:10.1103/PhysRevC.92.021301.
  18. Discovery and Assignment of Elements with Atomic Numbers 113, 115, 117 and 118. IUPAC (2015-12-30)
  19. Chatt, J. (1979). "Recommendations for the Naming of Elements of Atomic Numbers Greater than 100". Pure Appl. Chem. 51 (2): 381–384. doi:10.1351/pac197951020381.
  20. "IUPAC - International Union of Pure and Applied Chemistry: Discovery and Assignment of Elements with Atomic Numbers 113, 115, 117 and 118".
  21. Koppenol, W. H. (2002). "Naming of new elements (IUPAC Recommendations 2002)" (PDF). Pure and Applied Chemistry. 74 (5): 787. doi:10.1351/pac200274050787.
  22. "115-ый элемент Унунпентиум может появиться в таблице Менделеева". oane.ws (in Russian). 28 August 2013. Retrieved 23 September 2015. В свою очередь, российские физики предлагают свой вариант – ланжевений (Ln) в честь известного французского физика-теоретика прошлого столетия Ланжевена.
  23. Fedorova, Vera (30 March 2011). "Весенняя сессия Комитета полномочных представителей ОИЯИ". JINR (in Russian). Joint Institute for Nuclear Research. Retrieved 22 September 2015.
  24. Zavyalova, Victoria (25 August 2015). "Element 115, in Moscow's name". Russia & India Report. Retrieved 22 September 2015.
  25. 1 2 3 4 5 6 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.
  26. Considine, Glenn D.; Kulik, Peter H. (2002). Van Nostrand's scientific encyclopedia (9th ed.). Wiley-Interscience. ISBN 978-0-471-33230-5. OCLC 223349096.
  27. 1 2 Zagrebaev, V.; Greiner, W. (2008). "Synthesis of superheavy nuclei: A search for new production reactions". Physical Review C. 78 (3): 034610. arXiv:0807.2537Freely accessible. Bibcode:2008PhRvC..78c4610Z. doi:10.1103/PhysRevC.78.034610.
  28. "JINR Annual Reports 2000–2006". JINR. Retrieved 2013-08-27.
  29. 1 2 3 Thayer, John S. (2010). "Relativistic Effects and the Chemistry of the Heavier Main Group Elements". Relativistic Methods for Chemists. Springer. pp. 63–67, 83. doi:10.1007/978-1-4020-9975-5_2. ISBN 978-1-4020-9974-8.
  30. Faegri, K.; Saue, T. (2001). "Diatomic molecules between very heavy elements of group 13 and group 17: A study of relativistic effects on bonding". Journal of Chemical Physics. 115 (6): 2456. Bibcode:2001JChPh.115.2456F. doi:10.1063/1.1385366.
  31. Keller, O. L., Jr.; C. W. Nestor, Jr. (1974). "Predicted properties of the superheavy elements. III. Element 115, Eka-bismuth". Journal of Physical Chemistry. 78 (19): 1945. doi:10.1021/j100612a015.
  32. 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.
  33. 1 2 3 4 Eichler, Robert (2013). "First foot prints of chemistry on the shore of the Island of Superheavy Elements" (PDF). Journal of Physics: Conference Series. IOP Science. 420 (1): 012003. Bibcode:2013JPhCS.420a2003E. doi:10.1088/1742-6596/420/1/012003. Retrieved 11 September 2014.

External links

This article is issued from Wikipedia - version of the 12/4/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.