fbpx
Wikipedia

Copernicium

Copernicium is a synthetic chemical element; it has symbol Cn and atomic number 112. Its known isotopes are extremely radioactive, and have only been created in a laboratory. The most stable known isotope, copernicium-285, has a half-life of approximately 30 seconds. Copernicium was first created in 1996 by the GSI Helmholtz Centre for Heavy Ion Research near Darmstadt, Germany. It was named after the astronomer Nicolaus Copernicus.

Copernicium, 112Cn
Copernicium
Pronunciation/ˌkpərˈnɪsiəm/ (KOH-pər-NISS-ee-əm)
Mass number[285]
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
Groupgroup 12
Periodperiod 7
Block  d-block
Electron configuration[Rn] 5f14 6d10 7s2 (predicted)[1]
Electrons per shell2, 8, 18, 32, 32, 18, 2 (predicted)
Physical properties
Phase at STPliquid (predicted)[2][3]
Melting point283 ± 11 K ​(10 ± 11 °C, ​50 ± 20 °F) (predicted)[3]
Boiling point340 ± 10 K ​(67 ± 10 °C, ​153 ± 18 °F)[3] (predicted)
Density (near r.t.)14.0 g/cm3 (predicted)[3]
Triple point283 K, ​25 kPa (predicted)[3]
Atomic properties
Oxidation states0, (+1), +2, (+4), (+6) (parenthesized: prediction)[1][4][5][6]
Ionization energies
  • 1st: 1155 kJ/mol
  • 2nd: 2170 kJ/mol
  • 3rd: 3160 kJ/mol
  • (more) (all estimated)[1]
Atomic radiuscalculated: 147 pm[1][5] (predicted)
Covalent radius122 pm (predicted)[7]
Other properties
Natural occurrencesynthetic
Crystal structurehexagonal close-packed (hcp)

(predicted)[3]
CAS Number54084-26-3
History
Namingafter Nicolaus Copernicus
DiscoveryGesellschaft für Schwerionenforschung (1996)
Isotopes of copernicium
Main isotopes[8] Decay
abun­dance half-life (t1/2) mode pro­duct
283Cn synth 3.81 s[9] α96% 279Ds
SF4%
ε? 283Rg
285Cn synth 30 s α 281Ds
286Cn synth 8.4 s? SF
 Category: Copernicium
| references

In the periodic table of the elements, copernicium is a d-block transactinide element and a group 12 element. During reactions with gold, it has been shown[10] to be an extremely volatile element, so much so that it is possibly a gas or a volatile liquid 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 give up its 6d electrons instead of its 7s ones, and it may have more similarities to the noble gases such as radon rather than its group 12 homologues. Calculations indicate that copernicium may 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. Predictions vary on whether solid copernicium would be a metal, semiconductor, or insulator. Copernicium is one of the heaviest elements whose chemical properties have been experimentally investigated.

Introduction edit

Synthesis of superheavy nuclei edit

 
A graphic depiction of a nuclear fusion reaction. Two nuclei fuse into one, emitting a neutron. Reactions that created new elements to this moment were similar, with the only possible difference that several singular neutrons sometimes were released, or none at all.

A superheavy[a] atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size[b] into one; roughly, the more unequal the two nuclei in terms of mass, the greater the possibility that the two react.[16] The material made of the heavier nuclei is made into a target, which is then bombarded by the beam of lighter nuclei. Two nuclei can only fuse into one if they approach each other closely enough; normally, nuclei (all positively charged) repel each other due to electrostatic repulsion. The strong interaction can overcome this repulsion but only within a very short distance from a nucleus; beam nuclei are thus greatly accelerated in order to make such repulsion insignificant compared to the velocity of the beam nucleus.[17] The energy applied to the beam nuclei to accelerate them can cause them to reach speeds as high as one-tenth of the speed of light. However, if too much energy is applied, the beam nucleus can fall apart.[17]

Coming close enough alone is not enough for two nuclei to fuse: when two nuclei approach each other, they usually remain together for approximately 10−20 seconds and then part ways (not necessarily in the same composition as before the reaction) rather than form a single nucleus.[17][18] This happens because during the attempted formation of a single nucleus, electrostatic repulsion tears apart the nucleus that is being formed.[17] Each pair of a target and a beam is characterized by its cross section—the probability that fusion will occur if two nuclei approach one another expressed in terms of the transverse area that the incident particle must hit in order for the fusion to occur.[c] This fusion may occur as a result of the quantum effect in which nuclei can tunnel through electrostatic repulsion. If the two nuclei can stay close for past that phase, multiple nuclear interactions result in redistribution of energy and an energy equilibrium.[17]

External videos
  Visualization of unsuccessful nuclear fusion, based on calculations from the Australian National University[20]

The resulting merger is an excited state[21]—termed a compound nucleus—and thus it is very unstable.[17] To reach a more stable state, the temporary merger may fission without formation of a more stable nucleus.[22] Alternatively, the compound nucleus may eject a few neutrons, which would carry away the excitation energy; if the latter is not sufficient for a neutron expulsion, the merger would produce a gamma ray. This happens in approximately 10−16 seconds after the initial nuclear collision and results in creation of a more stable nucleus.[22] The definition by the IUPAC/IUPAP Joint Working Party (JWP) states that a chemical element can only be recognized as discovered if a nucleus of it has not decayed within 10−14 seconds. This value was chosen as an estimate of how long it takes a nucleus to acquire its outer electrons and thus display its chemical properties.[23][d]

Decay and detection edit

The beam passes through the target and reaches the next chamber, the separator; if a new nucleus is produced, it is carried with this beam.[25] In the separator, the newly produced nucleus is separated from other nuclides (that of the original beam and any other reaction products)[e] and transferred to a surface-barrier detector, which stops the nucleus. The exact location of the upcoming impact on the detector is marked; also marked are its energy and the time of the arrival.[25] The transfer takes about 10−6 seconds; in order to be detected, the nucleus must survive this long.[28] The nucleus is recorded again once its decay is registered, and the location, the energy, and the time of the decay are measured.[25]

Stability of a nucleus is provided by the strong interaction. However, its range is very short; as nuclei become larger, its influence on the outermost nucleons (protons and neutrons) weakens. At the same time, the nucleus is torn apart by electrostatic repulsion between protons, and its range is not limited.[29] Total binding energy provided by the strong interaction increases linearly with the number of nucleons, whereas electrostatic repulsion increases with the square of the atomic number, i.e. the latter grows faster and becomes increasingly important for heavy and superheavy nuclei.[30][31] Superheavy nuclei are thus theoretically predicted[32] and have so far been observed[33] to predominantly decay via decay modes that are caused by such repulsion: alpha decay and spontaneous fission.[f] Almost all alpha emitters have over 210 nucleons,[35] and the lightest nuclide primarily undergoing spontaneous fission has 238.[36] In both decay modes, nuclei are inhibited from decaying by corresponding energy barriers for each mode, but they can be tunnelled through.[30][31]

 
Scheme of an apparatus for creation of superheavy elements, based on the Dubna Gas-Filled Recoil Separator set up in the Flerov Laboratory of Nuclear Reactions in JINR. The trajectory within the detector and the beam focusing apparatus changes because of a dipole magnet in the former and quadrupole magnets in the latter.[37]

Alpha particles are commonly produced in radioactive decays because mass of an alpha particle per nucleon is small enough to leave some energy for the alpha particle to be used as kinetic energy to leave the nucleus.[38] Spontaneous fission is caused by electrostatic repulsion tearing the nucleus apart and produces various nuclei in different instances of identical nuclei fissioning.[31] As the atomic number increases, spontaneous fission rapidly becomes more important: spontaneous fission partial half-lives decrease by 23 orders of magnitude from uranium (element 92) to nobelium (element 102),[39] and by 30 orders of magnitude from thorium (element 90) to fermium (element 100).[40] The earlier liquid drop model thus suggested that spontaneous fission would occur nearly instantly due to disappearance of the fission barrier for nuclei with about 280 nucleons.[31][41] The later nuclear shell model suggested that nuclei with about 300 nucleons would form an island of stability in which nuclei will be more resistant to spontaneous fission and will primarily undergo alpha decay with longer half-lives.[31][41] Subsequent discoveries suggested that the predicted island might be further than originally anticipated; they also showed that nuclei intermediate between the long-lived actinides and the predicted island are deformed, and gain additional stability from shell effects.[42] Experiments on lighter superheavy nuclei,[43] as well as those closer to the expected island,[39] have shown greater than previously anticipated stability against spontaneous fission, showing the importance of shell effects on nuclei.[g]

Alpha decays are registered by the emitted alpha particles, and the decay products are easy to determine before the actual decay; if such a decay or a series of consecutive decays produces a known nucleus, the original product of a reaction can be easily determined.[h] (That all decays within a decay chain were indeed related to each other is established by the location of these decays, which must be in the same place.)[25] The known nucleus can be recognized by the specific characteristics of decay it undergoes such as decay energy (or more specifically, the kinetic energy of the emitted particle).[i] Spontaneous fission, however, produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[j]

The information available to physicists aiming to synthesize a superheavy element is thus the information collected at the detectors: location, energy, and time of arrival of a particle to the detector, and those of its decay. The physicists analyze this data and seek to conclude that it was indeed caused by a new element and could not have been caused by a different nuclide than the one claimed. Often, provided data is insufficient for a conclusion that a new element was definitely created and there is no other explanation for the observed effects; errors in interpreting data have been made.[k]

History edit

Discovery edit

Copernicium was first created on February 9, 1996, at the Gesellschaft für Schwerionenforschung (GSI) in Darmstadt, Germany, by Sigurd Hofmann, Victor Ninov et al.[54] 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 of copernicium was produced with a mass number of 277. (A second was originally reported, but was found to have been based on data fabricated by Ninov, and was thus retracted.)[54]

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.[55] 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.[56][57] 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.[58]

The IUPAC/IUPAP Joint Working Party (JWP) assessed the claim of copernicium's discovery by the GSI team in 2001[59] and 2003.[60] 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,[61] 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.[62] This decision was based on the confirmation of the decay properties of daughter nuclei as well as the confirmatory experiments at RIKEN.[63]

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),[63] 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 4-second half-life associated with 283Cn. (Another possibility is assignment to a metastable isomeric state, 283mCn.)[64] 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.[63]

Naming edit

 
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),[65] 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.[63][66] 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".[67]

During the standard six-month discussion period among the scientific community about the naming,[68][69] it was pointed out that the symbol Cp was previously associated with the name cassiopeium (cassiopium), now known as lutetium (Lu).[70][71] Moreover, Cp is frequently used today to mean the cyclopentadienyl ligand (C5H5).[72] Primarily because cassiopeium (Cp) was (until 1949) accepted by IUPAC as an alternative allowed name for lutetium,[73] 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.[68][74]

Isotopes edit

List of copernicium isotopes
Isotope Half-life[l] Decay
mode
Discovery
year
Discovery
reaction
Value ref
277Cn 0.79 ms [8] α 1996 208Pb(70Zn,n)
281Cn 0.18 s [75] α 2010 285Fl(—,α)
282Cn 0.83 ms [9] SF 2003 290Lv(—,2α)
283Cn 3.81 s [9] α, SF, EC? 2003 287Fl(—,α)
284Cn 121 ms [76] α, SF 2004 288Fl(—,α)
285Cn 30 s [8] α 1999 289Fl(—,α)
285mCn[m] 15 s [8] α 2012 293mLv(—,2α)
286Cn[m] 8.45 s [77] SF 2016 294Lv(—,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 mass numbers 277 and 281–286, and one unconfirmed metastable isomer in 285Cn has been reported.[78] Most of these decay predominantly through alpha decay, but some undergo spontaneous fission, and copernicium-283 may have an electron capture branch.[79]

The isotope copernicium-283 was instrumental in the confirmation of the discoveries of the elements flerovium and livermorium.[80]

Half-lives edit

All confirmed copernicium isotopes are extremely unstable and radioactive; in general, heavier isotopes are more stable than the lighter. The most stable known isotope, 285Cn, has a half-life of 30 seconds; 283Cn has a half-life of 4 seconds, and the unconfirmed 285mCn and 286Cn have half-lives of about 15 and 8.45 seconds respectively. Other isotopes have half-lives shorter than one second. 281Cn and 284Cn both have half-lives on the order of 0.1 seconds, and the other two isotopes have half-lives slightly under one millisecond.[79] It is predicted that the heavy isotopes 291Cn and 293Cn may have half-lives longer than a few decades, for they are predicted to lie near the center of the theoretical island of stability, 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.[81]

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.[79]

In 1999, American scientists at the University of California, Berkeley, announced that they had succeeded in synthesizing three atoms of 293Og.[82] These parent nuclei were reported to have successively emitted three alpha particles to form copernicium-281 nuclei, which were claimed to have undergone alpha decay, emitting alpha particles with decay energy 10.68 MeV and half-life 0.90 ms, but their claim was retracted in 2001[83] as it had been based on data fabricated by Ninov.[84] This isotope was truly produced in 2010 by the same team; the new data contradicted the previous fabricated data.[85]

Predicted properties edit

Very few properties of copernicium or its compounds have been measured; this is due to its extremely limited and expensive production[86] and the fact that copernicium (and its parents) decays very quickly. A few singular chemical properties have been measured, as well as the boiling point, but properties of the copernicium metal remain generally unknown and for the most part, only predictions are available.

Chemical edit

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. A standard reduction potential of +2.1 V is predicted for the Cn2+/Cn couple. Copernicium's predicted first ionization energy of 1155 kJ/mol almost matches that of the noble gas xenon at 1170.4 kJ/mol.[1] Copernicium's 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][87] 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] In opposition to the earlier suggestion,[88] ab initio calculations at the high level of accuracy[89] predicted that the chemistry of singly-valent copernicium resembles that of mercury rather than that of the noble gases. The latter result can be explained by the huge spin–orbit interaction which significantly lowers the energy of the vacant 7p1/2 state of copernicium.

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. As the most electronegative reactive element, fluorine may be the only element able to oxidise copernicium even further to the +4 and even +6 oxidation states in CnF4 and CnF6; the latter may require matrix-isolation conditions to be detected, as in the disputed detection of HgF4. CnF4 should be more stable than CnF2.[6] 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] The formation of thermodynamically stable copernicium(II) and (IV) fluorides would be analogous to the chemistry of xenon.[3] Analogous to mercury(II) cyanide (Hg(CN)2), copernicium is expected to form a stable cyanide, Cn(CN)2.[90]

Physical and atomic edit

Copernicium should be a dense metal, with a density of 14.0 g/cm3 in the liquid state at 300 K; this is similar to the known density of mercury, which is 13.534 g/cm3. (Solid copernicium at the same temperature should have a higher density of 14.7 g/cm3.) This results from the effects of copernicium's higher atomic weight being cancelled out by its larger interatomic distances compared to mercury.[3] Some calculations predicted copernicium to be a gas at room temperature due to its closed-shell electron configuration,[91] which would make it the first gaseous metal in the periodic table.[1][87] A 2019 calculation agrees with these predictions on the role of relativistic effects, suggesting that copernicium will be a volatile liquid bound by dispersion forces under standard conditions. Its melting point is estimated at 283±11 K and its boiling point at 340±10 K, the latter in agreement with the experimentally estimated value of 357+112
−108
 K
.[3] 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. Predictions of the expected band structure of copernicium are varied. Calculations in 2007 expected that copernicium may be a semiconductor[92] with a band gap of around 0.2 eV,[93] crystallizing in the hexagonal close-packed crystal structure.[93] 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.[94][95] 2019 calculations then suggested that in fact copernicium has a large band gap of 6.4 ± 0.2 eV, which should be similar to that of the noble gas radon (predicted as 7.1 eV) and would make it an insulator; bulk copernicium is predicted by these calculations to be bound mostly by dispersion forces, like the noble gases.[3] Like mercury, radon, and flerovium, but not oganesson (eka-radon), copernicium is calculated to have no electron affinity.[96]

Experimental atomic gas phase chemistry edit

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 is expected to have the ground state electron configuration [Rn] 5f14 6d10 7s2 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.[97]

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.[97] 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.[97] (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.)[98] In this experiment, two atoms of copernicium-283 were unambiguously identified and the adsorption properties were interpret to show that copernicium is a more volatile homologue of mercury, due to formation of a weak metal-metal bond with gold.[97] This agrees with general indications from some relativistic calculations that copernicium is "more or less" homologous to mercury.[99] However, it was pointed out in 2019 that this result may simply be due to strong dispersion interactions.[3]

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 in agreement with being the heaviest member of group 12.[97] 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.[92]

Because the lighter group 12 elements often occur as chalcogenide ores, experiments were conducted in 2015 to deposit copernicium atoms on a selenium surface to form copernicium selenide, CnSe. Reaction of copernicium atoms with trigonal selenium to form a selenide was 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.[100]

See also edit

Notes edit

  1. ^ In nuclear physics, an element is called heavy if its atomic number is high; lead (element 82) is one example of such a heavy element. The term "superheavy elements" typically refers to elements with atomic number greater than 103 (although there are other definitions, such as atomic number greater than 100[11] or 112;[12] sometimes, the term is presented an equivalent to the term "transactinide", which puts an upper limit before the beginning of the hypothetical superactinide series).[13] Terms "heavy isotopes" (of a given element) and "heavy nuclei" mean what could be understood in the common language—isotopes of high mass (for the given element) and nuclei of high mass, respectively.
  2. ^ In 2009, a team at the JINR led by Oganessian published results of their attempt to create hassium in a symmetric 136Xe + 136Xe reaction. They failed to observe a single atom in such a reaction, putting the upper limit on the cross section, the measure of probability of a nuclear reaction, as 2.5 pb.[14] In comparison, the reaction that resulted in hassium discovery, 208Pb + 58Fe, had a cross section of ~20 pb (more specifically, 19+19
    -11
     pb), as estimated by the discoverers.[15]
  3. ^ The amount of energy applied to the beam particle to accelerate it can also influence the value of cross section. For example, in the 28
    14
    Si
    + 1
    0
    n
    28
    13
    Al
    + 1
    1
    p
    reaction, cross section changes smoothly from 370 mb at 12.3 MeV to 160 mb at 18.3 MeV, with a broad peak at 13.5 MeV with the maximum value of 380 mb.[19]
  4. ^ This figure also marks the generally accepted upper limit for lifetime of a compound nucleus.[24]
  5. ^ This separation is based on that the resulting nuclei move past the target more slowly then the unreacted beam nuclei. The separator contains electric and magnetic fields whose effects on a moving particle cancel out for a specific velocity of a particle.[26] Such separation can also be aided by a time-of-flight measurement and a recoil energy measurement; a combination of the two may allow to estimate the mass of a nucleus.[27]
  6. ^ Not all decay modes are caused by electrostatic repulsion. For example, beta decay is caused by the weak interaction.[34]
  7. ^ It was already known by the 1960s that ground states of nuclei differed in energy and shape as well as that certain magic numbers of nucleons corresponded to greater stability of a nucleus. However, it was assumed that there was no nuclear structure in superheavy nuclei as they were too deformed to form one.[39]
  8. ^ Since mass of a nucleus is not measured directly but is rather calculated from that of another nucleus, such measurement is called indirect. Direct measurements are also possible, but for the most part they have remained unavailable for superheavy nuclei.[44] The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[45] Mass was determined from the location of a nucleus after the transfer (the location helps determine its trajectory, which is linked to the mass-to-charge ratio of the nucleus, since the transfer was done in presence of a magnet).[46]
  9. ^ If the decay occurred in a vacuum, then since total momentum of an isolated system before and after the decay must be preserved, the daughter nucleus would also receive a small velocity. The ratio of the two velocities, and accordingly the ratio of the kinetic energies, would thus be inverse to the ratio of the two masses. The decay energy equals the sum of the known kinetic energy of the alpha particle and that of the daughter nucleus (an exact fraction of the former).[35] The calculations hold for an experiment as well, but the difference is that the nucleus does not move after the decay because it is tied to the detector.
  10. ^ Spontaneous fission was discovered by Soviet physicist Georgy Flerov,[47] a leading scientist at JINR, and thus it was a "hobbyhorse" for the facility.[48] In contrast, the LBL scientists believed fission information was not sufficient for a claim of synthesis of an element. They believed spontaneous fission had not been studied enough to use it for identification of a new element, since there was a difficulty of establishing that a compound nucleus had only ejected neutrons and not charged particles like protons or alpha particles.[24] They thus preferred to link new isotopes to the already known ones by successive alpha decays.[47]
  11. ^ For instance, element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm, Stockholm County, Sweden.[49] There were no earlier definitive claims of creation of this element, and the element was assigned a name by its Swedish, American, and British discoverers, nobelium. It was later shown that the identification was incorrect.[50] The following year, RL was unable to reproduce the Swedish results and announced instead their synthesis of the element; that claim was also disproved later.[50] JINR insisted that they were the first to create the element and suggested a name of their own for the new element, joliotium;[51] the Soviet name was also not accepted (JINR later referred to the naming of the element 102 as "hasty").[52] This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements, signed 29 September 1992.[52] The name "nobelium" remained unchanged on account of its widespread usage.[53]
  12. ^ Different sources give different values for half-lives; the most recently published values are listed.
  13. ^ a b This isotope is unconfirmed

References edit

  1. ^ a b c d e f g h i j k l m n 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. ^ 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. ^ a b c d e f g h i j k Mewes, J.-M.; Smits, O. R.; Kresse, G.; Schwerdtfeger, P. (2019). "Copernicium is a Relativistic Noble Liquid". Angewandte Chemie International Edition. doi:10.1002/anie.201906966.
  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. doi:10.1007/978-3-642-37466-1_8. ISBN 978-3-642-37465-4. Retrieved April 21, 2018.
  5. ^ 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 October 4, 2013.
  6. ^ a b Hu, Shu-Xian; Zou, Wenli (September 23, 2021). "Stable copernicium hexafluoride (CnF6) with an oxidation state of VI+". Physical Chemistry Chemical Physics. 2022 (24): 321–325. doi:10.1039/D1CP04360A. PMID 34889909.
  7. ^ Chemical Data. Copernicium - Cn, Royal Chemical Society
  8. ^ a b c d 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.
  9. ^ a b c Oganessian, Yu. Ts.; Utyonkov, V. K.; Ibadullayev, D.; et al. (2022). "Investigation of 48Ca-induced reactions with 242Pu and 238U targets at the JINR Superheavy Element Factory". Physical Review C. 106 (24612). Bibcode:2022PhRvC.106b4612O. doi:10.1103/PhysRevC.106.024612. S2CID 251759318.
  10. ^ 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. S2CID 4347419.
  11. ^ Krämer, K. (2016). "Explainer: superheavy elements". Chemistry World. Retrieved March 15, 2020.
  12. ^ . Lawrence Livermore National Laboratory. Archived from the original on September 11, 2015. Retrieved March 15, 2020.
  13. ^ Eliav, E.; Kaldor, U.; Borschevsky, A. (2018). "Electronic Structure of the Transactinide Atoms". In Scott, R. A. (ed.). Encyclopedia of Inorganic and Bioinorganic Chemistry. John Wiley & Sons. pp. 1–16. doi:10.1002/9781119951438.eibc2632. ISBN 978-1-119-95143-8. S2CID 127060181.
  14. ^ Oganessian, Yu. Ts.; Dmitriev, S. N.; Yeremin, A. V.; et al. (2009). "Attempt to produce the isotopes of element 108 in the fusion reaction 136Xe + 136Xe". Physical Review C. 79 (2): 024608. doi:10.1103/PhysRevC.79.024608. ISSN 0556-2813.
  15. ^ Münzenberg, G.; Armbruster, P.; Folger, H.; et al. (1984). (PDF). Zeitschrift für Physik A. 317 (2): 235–236. Bibcode:1984ZPhyA.317..235M. doi:10.1007/BF01421260. S2CID 123288075. Archived from the original (PDF) on June 7, 2015. Retrieved October 20, 2012.
  16. ^ Subramanian, S. (August 28, 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved January 18, 2020.
  17. ^ a b c d e f Ivanov, D. (2019). "Сверхтяжелые шаги в неизвестное" [Superheavy steps into the unknown]. nplus1.ru (in Russian). Retrieved February 2, 2020.
  18. ^ Hinde, D. (2017). "Something new and superheavy at the periodic table". The Conversation. Retrieved January 30, 2020.
  19. ^ Kern, B. D.; Thompson, W. E.; Ferguson, J. M. (1959). "Cross sections for some (n, p) and (n, α) reactions". Nuclear Physics. 10: 226–234. Bibcode:1959NucPh..10..226K. doi:10.1016/0029-5582(59)90211-1.
  20. ^ Wakhle, A.; Simenel, C.; Hinde, D. J.; et al. (2015). Simenel, C.; Gomes, P. R. S.; Hinde, D. J.; et al. (eds.). "Comparing Experimental and Theoretical Quasifission Mass Angle Distributions". European Physical Journal Web of Conferences. 86: 00061. Bibcode:2015EPJWC..8600061W. doi:10.1051/epjconf/20158600061. hdl:1885/148847. ISSN 2100-014X.
  21. ^ "Nuclear Reactions" (PDF). pp. 7–8. Retrieved January 27, 2020. Published as Loveland, W. D.; Morrissey, D. J.; Seaborg, G. T. (2005). "Nuclear Reactions". Modern Nuclear Chemistry. John Wiley & Sons, Inc. pp. 249–297. doi:10.1002/0471768626.ch10. ISBN 978-0-471-76862-3.
  22. ^ a b Krása, A. (2010). "Neutron Sources for ADS". Faculty of Nuclear Sciences and Physical Engineering. Czech Technical University in Prague: 4–8. S2CID 28796927.
  23. ^ Wapstra, A. H. (1991). "Criteria that must be satisfied for the discovery of a new chemical element to be recognized" (PDF). Pure and Applied Chemistry. 63 (6): 883. doi:10.1351/pac199163060879. ISSN 1365-3075. S2CID 95737691.
  24. ^ a b Hyde, E. K.; Hoffman, D. C.; Keller, O. L. (1987). "A History and Analysis of the Discovery of Elements 104 and 105". Radiochimica Acta. 42 (2): 67–68. doi:10.1524/ract.1987.42.2.57. ISSN 2193-3405. S2CID 99193729.
  25. ^ a b c d Chemistry World (2016). "How to Make Superheavy Elements and Finish the Periodic Table [Video]". Scientific American. Retrieved January 27, 2020.
  26. ^ Hoffman, Ghiorso & Seaborg 2000, p. 334.
  27. ^ Hoffman, Ghiorso & Seaborg 2000, p. 335.
  28. ^ Zagrebaev, Karpov & Greiner 2013, p. 3.
  29. ^ Beiser 2003, p. 432.
  30. ^ a b Pauli, N. (2019). "Alpha decay" (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part). Université libre de Bruxelles. Retrieved February 16, 2020.
  31. ^ a b c d e Pauli, N. (2019). "Nuclear fission" (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part). Université libre de Bruxelles. Retrieved February 16, 2020.
  32. ^ Staszczak, A.; Baran, A.; Nazarewicz, W. (2013). "Spontaneous fission modes and lifetimes of superheavy elements in the nuclear density functional theory". Physical Review C. 87 (2): 024320–1. arXiv:1208.1215. Bibcode:2013PhRvC..87b4320S. doi:10.1103/physrevc.87.024320. ISSN 0556-2813.
  33. ^ Audi et al. 2017, pp. 030001-129–030001-138.
  34. ^ Beiser 2003, p. 439.
  35. ^ a b Beiser 2003, p. 433.
  36. ^ Audi et al. 2017, p. 030001-125.
  37. ^ Aksenov, N. V.; Steinegger, P.; Abdullin, F. Sh.; et al. (2017). "On the volatility of nihonium (Nh, Z = 113)". The European Physical Journal A. 53 (7): 158. Bibcode:2017EPJA...53..158A. doi:10.1140/epja/i2017-12348-8. ISSN 1434-6001. S2CID 125849923.
  38. ^ Beiser 2003, p. 432–433.
  39. ^ a b c Oganessian, Yu. (2012). "Nuclei in the "Island of Stability" of Superheavy Elements". Journal of Physics: Conference Series. 337 (1): 012005-1–012005-6. Bibcode:2012JPhCS.337a2005O. doi:10.1088/1742-6596/337/1/012005. ISSN 1742-6596.
  40. ^ Moller, P.; Nix, J. R. (1994). Fission properties of the heaviest elements (PDF). Dai 2 Kai Hadoron Tataikei no Simulation Symposium, Tokai-mura, Ibaraki, Japan. University of North Texas. Retrieved February 16, 2020.
  41. ^ a b Oganessian, Yu. Ts. (2004). "Superheavy elements". Physics World. 17 (7): 25–29. doi:10.1088/2058-7058/17/7/31. Retrieved February 16, 2020.
  42. ^ Schädel, M. (2015). "Chemistry of the superheavy elements". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 373 (2037): 20140191. Bibcode:2015RSPTA.37340191S. doi:10.1098/rsta.2014.0191. ISSN 1364-503X. PMID 25666065.
  43. ^ Hulet, E. K. (1989). Biomodal spontaneous fission. 50th Anniversary of Nuclear Fission, Leningrad, USSR. Bibcode:1989nufi.rept...16H.
  44. ^ Oganessian, Yu. Ts.; Rykaczewski, K. P. (2015). "A beachhead on the island of stability". Physics Today. 68 (8): 32–38. Bibcode:2015PhT....68h..32O. doi:10.1063/PT.3.2880. ISSN 0031-9228. OSTI 1337838. S2CID 119531411.
  45. ^ Grant, A. (2018). "Weighing the heaviest elements". Physics Today. doi:10.1063/PT.6.1.20181113a. S2CID 239775403.
  46. ^ Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table". Chemical & Engineering News. Retrieved January 27, 2020.
  47. ^ a b Robinson, A. E. (2019). "The Transfermium Wars: Scientific Brawling and Name-Calling during the Cold War". Distillations. Retrieved February 22, 2020.
  48. ^ "Популярная библиотека химических элементов. Сиборгий (экавольфрам)" [Popular library of chemical elements. Seaborgium (eka-tungsten)]. n-t.ru (in Russian). Retrieved January 7, 2020. Reprinted from "Экавольфрам" [Eka-tungsten]. Популярная библиотека химических элементов. Серебро – Нильсборий и далее [Popular library of chemical elements. Silver through nielsbohrium and beyond] (in Russian). Nauka. 1977.
  49. ^ "Nobelium - Element information, properties and uses | Periodic Table". Royal Society of Chemistry. Retrieved March 1, 2020.
  50. ^ a b Kragh 2018, pp. 38–39.
  51. ^ Kragh 2018, p. 40.
  52. ^ a b Ghiorso, A.; Seaborg, G. T.; Oganessian, Yu. Ts.; et al. (1993). "Responses on the report 'Discovery of the Transfermium elements' followed by reply to the responses by Transfermium Working Group" (PDF). Pure and Applied Chemistry. 65 (8): 1815–1824. doi:10.1351/pac199365081815. S2CID 95069384. (PDF) from the original on November 25, 2013. Retrieved September 7, 2016.
  53. ^ Commission on Nomenclature of Inorganic Chemistry (1997). "Names and symbols of transfermium elements (IUPAC Recommendations 1997)" (PDF). Pure and Applied Chemistry. 69 (12): 2471–2474. doi:10.1351/pac199769122471.
  54. ^ a b Hofmann, S.; et al. (1996). "The new element 112". Zeitschrift für Physik A. 354 (1): 229–230. Bibcode:1996ZPhyA.354..229H. doi:10.1007/BF02769517. S2CID 119975957.
  55. ^ Hofmann, S.; et al. (2000). (PDF). European Physical Journal A. Gesellschaft für Schwerionenforschung. 14 (2): 147–157. Bibcode:2002EPJA...14..147H. doi:10.1140/epja/i2001-10119-x. S2CID 8773326. Archived from the original (PDF) on February 27, 2008. Retrieved March 2, 2008.
  56. ^ Morita, K. (2004). "Decay of an Isotope 277112 produced by 208Pb + 70Zn reaction". In Penionzhkevich, Yu. E.; Cherepanov, E. A. (eds.). Exotic Nuclei: Proceedings of the International Symposium. World Scientific. pp. 188–191. doi:10.1142/9789812701749_0027.
  57. ^ Sumita, Takayuki; Morimoto, Kouji; Kaji, Daiya; Haba, Hiromitsu; Ozeki, Kazutaka; Sakai, Ryutaro; Yoneda, Akira; Yoshida, Atsushi; Hasebe, Hiroo; Katori, Kenji; Sato, Nozomi; Wakabayashi, Yasuo; Mitsuoka, Shin-Ichi; Goto, Shin-Ichi; Murakami, Masashi; Kariya, Yoshiki; Tokanai, Fuyuki; Mayama, Keita; Takeyama, Mirei; Moriya, Toru; Ideguchi, Eiji; Yamaguchi, Takayuki; Kikunaga, Hidetoshi; Chiba, Junsei; Morita, Kosuke (2013). "New Result on the Production of277Cn by the208Pb +70Zn Reaction". Journal of the Physical Society of Japan. 82 (2): 024202. Bibcode:2013JPSJ...82b4202S. doi:10.7566/JPSJ.82.024202.
  58. ^ Popeko, Andrey G. (2016). (PDF). jinr.ru. Joint Institute for Nuclear Research. Archived from the original (PDF) on February 4, 2018. Retrieved February 4, 2018.
  59. ^ Karol, P. J.; Nakahara, H.; Petley, B. W.; Vogt, E. (2001). (PDF). Pure and Applied Chemistry. 73 (6): 959–967. doi:10.1351/pac200173060959. S2CID 97615948. Archived from the original (PDF) on March 9, 2018. Retrieved January 9, 2008.
  60. ^ Karol, P. J.; Nakahara, H.; Petley, B. W.; Vogt, E. (2003). (PDF). Pure and Applied Chemistry. 75 (10): 1061–1611. doi:10.1351/pac200375101601. S2CID 95920517. Archived from the original (PDF) on August 22, 2016. Retrieved January 9, 2008.
  61. ^ Dressler, R.; Türler, A. (2001). (PDF). Annual Report. Paul Scherrer Institute. Archived from the original (PDF) on July 7, 2011.
  62. ^ . Gesellschaft für Schwerionenforschung. June 10, 2009. Archived from the original on August 23, 2009. Retrieved April 14, 2012.
  63. ^ a b c d 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. S2CID 95703833.
  64. ^ 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. (eds.). Exotic Nuclei: EXON-2016 Proceedings of the International Symposium on Exotic Nuclei. Exotic Nuclei. pp. 155–164. ISBN 9789813226555.
  65. ^ 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.
  66. ^ "New Chemical Element in the Periodic Table". Science Daily. June 11, 2009.
  67. ^ . Gesellschaft für Schwerionenforschung. July 14, 2009. Archived from the original on July 18, 2009.
  68. ^ a b "New element named 'copernicium'". BBC News. July 16, 2009. Retrieved February 22, 2010.
  69. ^ . IUPAC. July 20, 2009. Archived from the original on November 27, 2012. Retrieved April 14, 2012.
  70. ^ Meija, Juris (2009). "The need for a fresh symbol to designate copernicium". Nature. 461 (7262): 341. Bibcode:2009Natur.461..341M. doi:10.1038/461341c. PMID 19759598.
  71. ^ van der Krogt, P. "Lutetium". Elementymology & Elements Multidict. Retrieved February 22, 2010.
  72. ^ "Minutes, Division VIII Committee meeting, Glasgow, 2009" (PDF). iupac.org. IUPAC. 2009. Retrieved January 11, 2024.
  73. ^ Tatsumi, Kazuyuki; Corish, John (2010). "Name and symbol of the element with atomic number 112 (IUPAC Recommendations 2010)" (PDF). Pure and Applied Chemistry. 82 (3): 753–755. doi:10.1351/PAC-REC-09-08-20. Retrieved January 11, 2024.
  74. ^ . IUPAC. February 19, 2010. Archived from the original on March 4, 2016. Retrieved April 13, 2012.
  75. ^ Utyonkov, V. K.; Brewer, N. T.; Oganessian, Yu. Ts.; et al. (January 30, 2018). "Neutron-deficient superheavy nuclei obtained in the 240Pu+48Ca reaction". Physical Review C. 97 (14320): 014320. Bibcode:2018PhRvC..97a4320U. doi:10.1103/PhysRevC.97.014320.
  76. ^ Såmark-Roth, A.; Cox, D. M.; Rudolph, D.; et al. (2021). "Spectroscopy along Flerovium Decay Chains: Discovery of 280Ds and an Excited State in 282Cn". Physical Review Letters. 126 (3): 032503. Bibcode:2021PhRvL.126c2503S. doi:10.1103/PhysRevLett.126.032503. hdl:10486/705608. PMID 33543956. S2CID 231818619.
  77. ^ Kaji, Daiya; Morita, Kosuke; Morimoto, Kouji; Haba, Hiromitsu; Asai, Masato; Fujita, Kunihiro; Gan, Zaiguo; Geissel, Hans; Hasebe, Hiroo; Hofmann, Sigurd; Huang, MingHui; Komori, Yukiko; Ma, Long; Maurer, Joachim; Murakami, Masashi; Takeyama, Mirei; Tokanai, Fuyuki; Tanaka, Taiki; Wakabayashi, Yasuo; Yamaguchi, Takayuki; Yamaki, Sayaka; Yoshida, Atsushi (2017). "Study of the Reaction 48Ca + 248Cm → 296Lv* at RIKEN-GARIS". Journal of the Physical Society of Japan. 86 (3): 034201–1–7. Bibcode:2017JPSJ...86c4201K. doi:10.7566/JPSJ.86.034201.
  78. ^ Hofmann, S.; Heinz, S.; Mann, R.; et al. (2012). "The reaction 48Ca + 248Cm → 296116* studied at the GSI-SHIP". The European Physical Journal A. 48 (5): 62. Bibcode:2012EPJA...48...62H. doi:10.1140/epja/i2012-12062-1. S2CID 121930293.
  79. ^ a b c Holden, N. E. (2004). "Table of the Isotopes". In D. R. Lide (ed.). CRC Handbook of Chemistry and Physics (85th ed.). CRC Press. Section 11. ISBN 978-0-8493-0485-9.
  80. ^ 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. S2CID 98065999.
  81. ^ Zagrebaev, Karpov & Greiner 2013, pp. 1–15.
  82. ^ Ninov, V.; et al. (1999). "Observation of Superheavy Nuclei Produced in the Reaction of 86
    Kr
    with 208
    Pb
    ". Physical Review Letters. 83 (6): 1104–1107. Bibcode:1999PhRvL..83.1104N. doi:10.1103/PhysRevLett.83.1104.
  83. ^ Public Affairs Department (July 21, 2001). . Berkeley Lab. Archived from the original on January 29, 2008. Retrieved January 18, 2008.
  84. ^ "At Lawrence Berkeley, Physicists Say a Colleague Took Them for a Ride" George Johnson, The New York Times, 15 October 2002
  85. ^ Public Affairs Department (October 26, 2010). "Six New Isotopes of the Superheavy Elements Discovered: Moving Closer to Understanding the Island of Stability". Berkeley Lab. Retrieved April 25, 2011.
  86. ^ Subramanian, S. (August 28, 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved January 18, 2020.
  87. ^ a b "Chemistry on the islands of stability", New Scientist, 11 September 1975, p. 574, ISSN 1032-1233
  88. ^ Pitzer, K. S. (1975). "Are elements 112, 114, and 118 relatively inert gases?". The Journal of Chemical Physics. 63 (2): 1032–1033. doi:10.1063/1.431398.
  89. ^ Mosyagin, N. S.; Isaev, T. A.; Titov, A. V. (2006). "Is E112 a relatively inert element? Benchmark relativistic correlation study of spectroscopic constants in E112H and its cation". The Journal of Chemical Physics. 124 (22): 224302. arXiv:physics/0508024. Bibcode:2006JChPh.124v4302M. doi:10.1063/1.2206189. PMID 16784269. S2CID 119339584.
  90. ^ Demissie, Taye B.; Ruud, Kenneth (February 25, 2017). "Darmstadtium, roentgenium, and copernicium form strong bonds with cyanide". International Journal of Quantum Chemistry. 2017: e25393. doi:10.1002/qua.25393. hdl:10037/13632.
  91. ^ Kratz, Jens Volker. The Impact of Superheavy Elements on the Chemical and Physical Sciences June 14, 2022, at the Wayback Machine. 4th International Conference on the Chemistry and Physics of the Transactinide Elements, 5–11 September 2011, Sochi, Russia
  92. ^ a b 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): 3262–3266. doi:10.1002/anie.200705019. PMID 18338360.
  93. ^ a b Gaston, Nicola; Opahle, Ingo; Gäggeler, Heinz W.; Schwerdtfeger, Peter (2007). "Is eka-mercury (element 112) a group 12 metal?". Angewandte Chemie. 46 (10): 1663–1666. doi:10.1002/anie.200604262. PMID 17397075. Retrieved November 5, 2013.
  94. ^ 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. Bibcode:2018SSCom.269...16G. doi:10.1016/j.ssc.2017.10.009. Retrieved March 28, 2018.
  95. ^ Čenčariková, Hana; Legut, Dominik (2018). "The effect of relativity on stability of Copernicium phases, their electronic structure and mechanical properties". Physica B. 536: 576–582. arXiv:1810.01955. Bibcode:2018PhyB..536..576C. doi:10.1016/j.physb.2017.11.035. S2CID 119100368.
  96. ^ Borschevsky, Anastasia; Pershina, Valeria; Kaldor, Uzi; Eliav, Ephraim. (PDF). www.kernchemie.uni-mainz.de. Johannes Gutenberg University Mainz. Archived from the original (PDF) on January 15, 2018. Retrieved January 15, 2018.
  97. ^ a b c d e Gäggeler, H. W. (2007). (PDF). Paul Scherrer Institute. pp. 26–28. Archived from the original (PDF) on February 20, 2012.
  98. ^ Moody, Ken (2013). "Synthesis of Superheavy Elements". In Schädel, Matthias; Shaughnessy, Dawn (eds.). The Chemistry of Superheavy Elements (2nd ed.). Springer Science & Business Media. pp. 24–28. ISBN 9783642374661.
  99. ^ 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 February 17, 2018.
  100. ^ "Annual Report 2015: Laboratory of Radiochemistry and Environmental Chemistry" (PDF). Paul Scherrer Institute. 2015. p. 3.

Bibliography edit

  • Audi, G.; Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S. (2017). "The NUBASE2016 evaluation of nuclear properties". Chinese Physics C. 41 (3). 030001. Bibcode:2017ChPhC..41c0001A. doi:10.1088/1674-1137/41/3/030001.
  • Beiser, A. (2003). Concepts of modern physics (6th ed.). McGraw-Hill. ISBN 978-0-07-244848-1. OCLC 48965418.
  • Hoffman, D. C.; Ghiorso, A.; Seaborg, G. T. (2000). The Transuranium People: The Inside Story. World Scientific. ISBN 978-1-78-326244-1.
  • Kragh, H. (2018). From Transuranic to Superheavy Elements: A Story of Dispute and Creation. Springer. ISBN 978-3-319-75813-8.
  • Zagrebaev, Valeriy; Karpov, Alexander; Greiner, Walter (2013). "Future of superheavy element research: Which nuclei could be synthesized within the next few years?" (PDF). 11th International Conference on Nucleus-Nucleus Collisions (NN2012). Journal of Physics: Conference Series. Vol. 420. IOP Publishing. doi:10.1088/1742-6596/420/1/012001. Retrieved August 20, 2013.

External links edit

copernicium, redirects, here, fictional, character, dragon, ball, synthetic, chemical, element, symbol, atomic, number, known, isotopes, extremely, radioactive, have, only, been, created, laboratory, most, stable, known, isotope, copernicium, half, life, appro. Uub redirects here For the fictional character see Uub Dragon Ball Copernicium is a synthetic chemical element it has symbol Cn and atomic number 112 Its known isotopes are extremely radioactive and have only been created in a laboratory The most stable known isotope copernicium 285 has a half life of approximately 30 seconds Copernicium was first created in 1996 by the GSI Helmholtz Centre for Heavy Ion Research near Darmstadt Germany It was named after the astronomer Nicolaus Copernicus Copernicium 112CnCoperniciumPronunciation ˌ k oʊ p er ˈ n ɪ s i e m wbr KOH per NISS ee em Mass number 285 Copernicium in the periodic tableHydrogen HeliumLithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine NeonSodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine ArgonPotassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine KryptonRubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine XenonCaesium 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 RadonFrancium 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 roentgenium copernicium nihoniumAtomic number Z 112Groupgroup 12Periodperiod 7Block d blockElectron configuration Rn 5f14 6d10 7s2 predicted 1 Electrons per shell2 8 18 32 32 18 2 predicted Physical propertiesPhase at STPliquid predicted 2 3 Melting point283 11 K 10 11 C 50 20 F predicted 3 Boiling point340 10 K 67 10 C 153 18 F 3 predicted Density near r t 14 0 g cm3 predicted 3 Triple point283 K 25 kPa predicted 3 Atomic propertiesOxidation states0 1 2 4 6 parenthesized prediction 1 4 5 6 Ionization energies1st 1155 kJ mol2nd 2170 kJ mol3rd 3160 kJ mol more all estimated 1 Atomic radiuscalculated 147 pm 1 5 predicted Covalent radius122 pm predicted 7 Other propertiesNatural occurrencesyntheticCrystal structure hexagonal close packed hcp predicted 3 CAS Number54084 26 3HistoryNamingafter Nicolaus CopernicusDiscoveryGesellschaft fur Schwerionenforschung 1996 Isotopes of coperniciumveMain isotopes 8 Decayabun dance half life t1 2 mode pro duct283Cn synth 3 81 s 9 a 96 279DsSF 4 e 283Rg285Cn synth 30 s a 281Ds286Cn synth 8 4 s SF Category Coperniciumviewtalkedit referencesIn the periodic table of the elements copernicium is a d block transactinide element and a group 12 element During reactions with gold it has been shown 10 to be an extremely volatile element so much so that it is possibly a gas or a volatile liquid 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 give up its 6d electrons instead of its 7s ones and it may have more similarities to the noble gases such as radon rather than its group 12 homologues Calculations indicate that copernicium may 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 Predictions vary on whether solid copernicium would be a metal semiconductor or insulator Copernicium is one of the heaviest elements whose chemical properties have been experimentally investigated Contents 1 Introduction 1 1 Synthesis of superheavy nuclei 1 2 Decay and detection 2 History 2 1 Discovery 2 2 Naming 3 Isotopes 3 1 Half lives 4 Predicted properties 4 1 Chemical 4 2 Physical and atomic 5 Experimental atomic gas phase chemistry 6 See also 7 Notes 8 References 9 Bibliography 10 External linksIntroduction editThis section is an excerpt from Superheavy element Introduction edit Synthesis of superheavy nuclei edit nbsp A graphic depiction of a nuclear fusion reaction Two nuclei fuse into one emitting a neutron Reactions that created new elements to this moment were similar with the only possible difference that several singular neutrons sometimes were released or none at all A superheavy a atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size b into one roughly the more unequal the two nuclei in terms of mass the greater the possibility that the two react 16 The material made of the heavier nuclei is made into a target which is then bombarded by the beam of lighter nuclei Two nuclei can only fuse into one if they approach each other closely enough normally nuclei all positively charged repel each other due to electrostatic repulsion The strong interaction can overcome this repulsion but only within a very short distance from a nucleus beam nuclei are thus greatly accelerated in order to make such repulsion insignificant compared to the velocity of the beam nucleus 17 The energy applied to the beam nuclei to accelerate them can cause them to reach speeds as high as one tenth of the speed of light However if too much energy is applied the beam nucleus can fall apart 17 Coming close enough alone is not enough for two nuclei to fuse when two nuclei approach each other they usually remain together for approximately 10 20 seconds and then part ways not necessarily in the same composition as before the reaction rather than form a single nucleus 17 18 This happens because during the attempted formation of a single nucleus electrostatic repulsion tears apart the nucleus that is being formed 17 Each pair of a target and a beam is characterized by its cross section the probability that fusion will occur if two nuclei approach one another expressed in terms of the transverse area that the incident particle must hit in order for the fusion to occur c This fusion may occur as a result of the quantum effect in which nuclei can tunnel through electrostatic repulsion If the two nuclei can stay close for past that phase multiple nuclear interactions result in redistribution of energy and an energy equilibrium 17 External videos nbsp Visualization of unsuccessful nuclear fusion based on calculations from the Australian National University 20 The resulting merger is an excited state 21 termed a compound nucleus and thus it is very unstable 17 To reach a more stable state the temporary merger may fission without formation of a more stable nucleus 22 Alternatively the compound nucleus may eject a few neutrons which would carry away the excitation energy if the latter is not sufficient for a neutron expulsion the merger would produce a gamma ray This happens in approximately 10 16 seconds after the initial nuclear collision and results in creation of a more stable nucleus 22 The definition by the IUPAC IUPAP Joint Working Party JWP states that a chemical element can only be recognized as discovered if a nucleus of it has not decayed within 10 14 seconds This value was chosen as an estimate of how long it takes a nucleus to acquire its outer electrons and thus display its chemical properties 23 d Decay and detection edit The beam passes through the target and reaches the next chamber the separator if a new nucleus is produced it is carried with this beam 25 In the separator the newly produced nucleus is separated from other nuclides that of the original beam and any other reaction products e and transferred to a surface barrier detector which stops the nucleus The exact location of the upcoming impact on the detector is marked also marked are its energy and the time of the arrival 25 The transfer takes about 10 6 seconds in order to be detected the nucleus must survive this long 28 The nucleus is recorded again once its decay is registered and the location the energy and the time of the decay are measured 25 Stability of a nucleus is provided by the strong interaction However its range is very short as nuclei become larger its influence on the outermost nucleons protons and neutrons weakens At the same time the nucleus is torn apart by electrostatic repulsion between protons and its range is not limited 29 Total binding energy provided by the strong interaction increases linearly with the number of nucleons whereas electrostatic repulsion increases with the square of the atomic number i e the latter grows faster and becomes increasingly important for heavy and superheavy nuclei 30 31 Superheavy nuclei are thus theoretically predicted 32 and have so far been observed 33 to predominantly decay via decay modes that are caused by such repulsion alpha decay and spontaneous fission f Almost all alpha emitters have over 210 nucleons 35 and the lightest nuclide primarily undergoing spontaneous fission has 238 36 In both decay modes nuclei are inhibited from decaying by corresponding energy barriers for each mode but they can be tunnelled through 30 31 nbsp Scheme of an apparatus for creation of superheavy elements based on the Dubna Gas Filled Recoil Separator set up in the Flerov Laboratory of Nuclear Reactions in JINR The trajectory within the detector and the beam focusing apparatus changes because of a dipole magnet in the former and quadrupole magnets in the latter 37 Alpha particles are commonly produced in radioactive decays because mass of an alpha particle per nucleon is small enough to leave some energy for the alpha particle to be used as kinetic energy to leave the nucleus 38 Spontaneous fission is caused by electrostatic repulsion tearing the nucleus apart and produces various nuclei in different instances of identical nuclei fissioning 31 As the atomic number increases spontaneous fission rapidly becomes more important spontaneous fission partial half lives decrease by 23 orders of magnitude from uranium element 92 to nobelium element 102 39 and by 30 orders of magnitude from thorium element 90 to fermium element 100 40 The earlier liquid drop model thus suggested that spontaneous fission would occur nearly instantly due to disappearance of the fission barrier for nuclei with about 280 nucleons 31 41 The later nuclear shell model suggested that nuclei with about 300 nucleons would form an island of stability in which nuclei will be more resistant to spontaneous fission and will primarily undergo alpha decay with longer half lives 31 41 Subsequent discoveries suggested that the predicted island might be further than originally anticipated they also showed that nuclei intermediate between the long lived actinides and the predicted island are deformed and gain additional stability from shell effects 42 Experiments on lighter superheavy nuclei 43 as well as those closer to the expected island 39 have shown greater than previously anticipated stability against spontaneous fission showing the importance of shell effects on nuclei g Alpha decays are registered by the emitted alpha particles and the decay products are easy to determine before the actual decay if such a decay or a series of consecutive decays produces a known nucleus the original product of a reaction can be easily determined h That all decays within a decay chain were indeed related to each other is established by the location of these decays which must be in the same place 25 The known nucleus can be recognized by the specific characteristics of decay it undergoes such as decay energy or more specifically the kinetic energy of the emitted particle i Spontaneous fission however produces various nuclei as products so the original nuclide cannot be determined from its daughters j The information available to physicists aiming to synthesize a superheavy element is thus the information collected at the detectors location energy and time of arrival of a particle to the detector and those of its decay The physicists analyze this data and seek to conclude that it was indeed caused by a new element and could not have been caused by a different nuclide than the one claimed Often provided data is insufficient for a conclusion that a new element was definitely created and there is no other explanation for the observed effects errors in interpreting data have been made k History editDiscovery edit Copernicium was first created on February 9 1996 at the Gesellschaft fur Schwerionenforschung GSI in Darmstadt Germany by Sigurd Hofmann Victor Ninov et al 54 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 of copernicium was produced with a mass number of 277 A second was originally reported but was found to have been based on data fabricated by Ninov and was thus retracted 54 20882 Pb 7030 Zn 278112 Cn 277112 Cn 10 nIn May 2000 the GSI successfully repeated the experiment to synthesize a further atom of copernicium 277 55 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 56 57 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 58 The IUPAC IUPAP Joint Working Party JWP assessed the claim of copernicium s discovery by the GSI team in 2001 59 and 2003 60 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 61 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 62 This decision was based on the confirmation of the decay properties of daughter nuclei as well as the confirmatory experiments at RIKEN 63 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 63 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 4 second half life associated with 283Cn Another possibility is assignment to a metastable isomeric state 283mCn 64 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 63 Naming edit nbsp 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 65 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 63 66 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 67 During the standard six month discussion period among the scientific community about the naming 68 69 it was pointed out that the symbol Cp was previously associated with the name cassiopeium cassiopium now known as lutetium Lu 70 71 Moreover Cp is frequently used today to mean the cyclopentadienyl ligand C5H5 72 Primarily because cassiopeium Cp was until 1949 accepted by IUPAC as an alternative allowed name for lutetium 73 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 68 74 Isotopes editMain article Isotopes of copernicium List of copernicium isotopes vte Isotope Half life l Decaymode Discoveryyear DiscoveryreactionValue ref277Cn 0 79 ms 8 a 1996 208Pb 70Zn n 281Cn 0 18 s 75 a 2010 285Fl a 282Cn 0 83 ms 9 SF 2003 290Lv 2a 283Cn 3 81 s 9 a SF EC 2003 287Fl a 284Cn 121 ms 76 a SF 2004 288Fl a 285Cn 30 s 8 a 1999 289Fl a 285mCn m 15 s 8 a 2012 293mLv 2a 286Cn m 8 45 s 77 SF 2016 294Lv 2a 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 mass numbers 277 and 281 286 and one unconfirmed metastable isomer in 285Cn has been reported 78 Most of these decay predominantly through alpha decay but some undergo spontaneous fission and copernicium 283 may have an electron capture branch 79 The isotope copernicium 283 was instrumental in the confirmation of the discoveries of the elements flerovium and livermorium 80 Half lives edit All confirmed copernicium isotopes are extremely unstable and radioactive in general heavier isotopes are more stable than the lighter The most stable known isotope 285Cn has a half life of 30 seconds 283Cn has a half life of 4 seconds and the unconfirmed 285mCn and 286Cn have half lives of about 15 and 8 45 seconds respectively Other isotopes have half lives shorter than one second 281Cn and 284Cn both have half lives on the order of 0 1 seconds and the other two isotopes have half lives slightly under one millisecond 79 It is predicted that the heavy isotopes 291Cn and 293Cn may have half lives longer than a few decades for they are predicted to lie near the center of the theoretical island of stability 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 81 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 79 In 1999 American scientists at the University of California Berkeley announced that they had succeeded in synthesizing three atoms of 293Og 82 These parent nuclei were reported to have successively emitted three alpha particles to form copernicium 281 nuclei which were claimed to have undergone alpha decay emitting alpha particles with decay energy 10 68 MeV and half life 0 90 ms but their claim was retracted in 2001 83 as it had been based on data fabricated by Ninov 84 This isotope was truly produced in 2010 by the same team the new data contradicted the previous fabricated data 85 Predicted properties editVery few properties of copernicium or its compounds have been measured this is due to its extremely limited and expensive production 86 and the fact that copernicium and its parents decays very quickly A few singular chemical properties have been measured as well as the boiling point but properties of the copernicium metal remain generally unknown and for the most part only predictions are available Chemical edit 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 A standard reduction potential of 2 1 V is predicted for the Cn2 Cn couple Copernicium s predicted first ionization energy of 1155 kJ mol almost matches that of the noble gas xenon at 1170 4 kJ mol 1 Copernicium s 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 87 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 In opposition to the earlier suggestion 88 ab initio calculations at the high level of accuracy 89 predicted that the chemistry of singly valent copernicium resembles that of mercury rather than that of the noble gases The latter result can be explained by the huge spin orbit interaction which significantly lowers the energy of the vacant 7p1 2 state of copernicium 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 As the most electronegative reactive element fluorine may be the only element able to oxidise copernicium even further to the 4 and even 6 oxidation states in CnF4 and CnF6 the latter may require matrix isolation conditions to be detected as in the disputed detection of HgF4 CnF4 should be more stable than CnF2 6 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 The formation of thermodynamically stable copernicium II and IV fluorides would be analogous to the chemistry of xenon 3 Analogous to mercury II cyanide Hg CN 2 copernicium is expected to form a stable cyanide Cn CN 2 90 Physical and atomic edit Copernicium should be a dense metal with a density of 14 0 g cm3 in the liquid state at 300 K this is similar to the known density of mercury which is 13 534 g cm3 Solid copernicium at the same temperature should have a higher density of 14 7 g cm3 This results from the effects of copernicium s higher atomic weight being cancelled out by its larger interatomic distances compared to mercury 3 Some calculations predicted copernicium to be a gas at room temperature due to its closed shell electron configuration 91 which would make it the first gaseous metal in the periodic table 1 87 A 2019 calculation agrees with these predictions on the role of relativistic effects suggesting that copernicium will be a volatile liquid bound by dispersion forces under standard conditions Its melting point is estimated at 283 11 K and its boiling point at 340 10 K the latter in agreement with the experimentally estimated value of 357 112 108 K 3 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 Predictions of the expected band structure of copernicium are varied Calculations in 2007 expected that copernicium may be a semiconductor 92 with a band gap of around 0 2 eV 93 crystallizing in the hexagonal close packed crystal structure 93 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 94 95 2019 calculations then suggested that in fact copernicium has a large band gap of 6 4 0 2 eV which should be similar to that of the noble gas radon predicted as 7 1 eV and would make it an insulator bulk copernicium is predicted by these calculations to be bound mostly by dispersion forces like the noble gases 3 Like mercury radon and flerovium but not oganesson eka radon copernicium is calculated to have no electron affinity 96 Experimental atomic gas phase chemistry editInterest 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 is expected to have the ground state electron configuration Rn 5f14 6d10 7s2 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 97 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 97 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 97 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 98 In this experiment two atoms of copernicium 283 were unambiguously identified and the adsorption properties were interpret to show that copernicium is a more volatile homologue of mercury due to formation of a weak metal metal bond with gold 97 This agrees with general indications from some relativistic calculations that copernicium is more or less homologous to mercury 99 However it was pointed out in 2019 that this result may simply be due to strong dispersion interactions 3 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 in agreement with being the heaviest member of group 12 97 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 92 Because the lighter group 12 elements often occur as chalcogenide ores experiments were conducted in 2015 to deposit copernicium atoms on a selenium surface to form copernicium selenide CnSe Reaction of copernicium atoms with trigonal selenium to form a selenide was observed with DHadsCn t Se gt 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 100 See also editIsland of stabilityNotes edit In nuclear physics an element is called heavy if its atomic number is high lead element 82 is one example of such a heavy element The term superheavy elements typically refers to elements with atomic number greater than 103 although there are other definitions such as atomic number greater than 100 11 or 112 12 sometimes the term is presented an equivalent to the term transactinide which puts an upper limit before the beginning of the hypothetical superactinide series 13 Terms heavy isotopes of a given element and heavy nuclei mean what could be understood in the common language isotopes of high mass for the given element and nuclei of high mass respectively In 2009 a team at the JINR led by Oganessian published results of their attempt to create hassium in a symmetric 136Xe 136Xe reaction They failed to observe a single atom in such a reaction putting the upper limit on the cross section the measure of probability of a nuclear reaction as 2 5 pb 14 In comparison the reaction that resulted in hassium discovery 208Pb 58Fe had a cross section of 20 pb more specifically 19 19 11 pb as estimated by the discoverers 15 The amount of energy applied to the beam particle to accelerate it can also influence the value of cross section For example in the 2814 Si 10 n 2813 Al 11 p reaction cross section changes smoothly from 370 mb at 12 3 MeV to 160 mb at 18 3 MeV with a broad peak at 13 5 MeV with the maximum value of 380 mb 19 This figure also marks the generally accepted upper limit for lifetime of a compound nucleus 24 This separation is based on that the resulting nuclei move past the target more slowly then the unreacted beam nuclei The separator contains electric and magnetic fields whose effects on a moving particle cancel out for a specific velocity of a particle 26 Such separation can also be aided by a time of flight measurement and a recoil energy measurement a combination of the two may allow to estimate the mass of a nucleus 27 Not all decay modes are caused by electrostatic repulsion For example beta decay is caused by the weak interaction 34 It was already known by the 1960s that ground states of nuclei differed in energy and shape as well as that certain magic numbers of nucleons corresponded to greater stability of a nucleus However it was assumed that there was no nuclear structure in superheavy nuclei as they were too deformed to form one 39 Since mass of a nucleus is not measured directly but is rather calculated from that of another nucleus such measurement is called indirect Direct measurements are also possible but for the most part they have remained unavailable for superheavy nuclei 44 The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL 45 Mass was determined from the location of a nucleus after the transfer the location helps determine its trajectory which is linked to the mass to charge ratio of the nucleus since the transfer was done in presence of a magnet 46 If the decay occurred in a vacuum then since total momentum of an isolated system before and after the decay must be preserved the daughter nucleus would also receive a small velocity The ratio of the two velocities and accordingly the ratio of the kinetic energies would thus be inverse to the ratio of the two masses The decay energy equals the sum of the known kinetic energy of the alpha particle and that of the daughter nucleus an exact fraction of the former 35 The calculations hold for an experiment as well but the difference is that the nucleus does not move after the decay because it is tied to the detector Spontaneous fission was discovered by Soviet physicist Georgy Flerov 47 a leading scientist at JINR and thus it was a hobbyhorse for the facility 48 In contrast the LBL scientists believed fission information was not sufficient for a claim of synthesis of an element They believed spontaneous fission had not been studied enough to use it for identification of a new element since there was a difficulty of establishing that a compound nucleus had only ejected neutrons and not charged particles like protons or alpha particles 24 They thus preferred to link new isotopes to the already known ones by successive alpha decays 47 For instance element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm Stockholm County Sweden 49 There were no earlier definitive claims of creation of this element and the element was assigned a name by its Swedish American and British discoverers nobelium It was later shown that the identification was incorrect 50 The following year RL was unable to reproduce the Swedish results and announced instead their synthesis of the element that claim was also disproved later 50 JINR insisted that they were the first to create the element and suggested a name of their own for the new element joliotium 51 the Soviet name was also not accepted JINR later referred to the naming of the element 102 as hasty 52 This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements signed 29 September 1992 52 The name nobelium remained unchanged on account of its widespread usage 53 Different sources give different values for half lives the most recently published values are listed a b This isotope is unconfirmedReferences edit a b c d e f g h i j k l m n 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 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 a b c d e f g h i j k Mewes J M Smits O R Kresse G Schwerdtfeger P 2019 Copernicium is a Relativistic Noble Liquid Angewandte Chemie International Edition doi 10 1002 anie 201906966 Gaggeler Heinz W Turler Andreas 2013 Gas Phase Chemistry of Superheavy Elements The Chemistry of Superheavy Elements Springer Science Business Media pp 415 483 doi 10 1007 978 3 642 37466 1 8 ISBN 978 3 642 37465 4 Retrieved April 21 2018 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 October 4 2013 a b Hu Shu Xian Zou Wenli September 23 2021 Stable copernicium hexafluoride CnF6 with an oxidation state of VI Physical Chemistry Chemical Physics 2022 24 321 325 doi 10 1039 D1CP04360A PMID 34889909 Chemical Data Copernicium Cn Royal Chemical Society a b c d 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 a b c Oganessian Yu Ts Utyonkov V K Ibadullayev D et al 2022 Investigation of 48Ca induced reactions with 242Pu and 238U targets at the JINR Superheavy Element Factory Physical Review C 106 24612 Bibcode 2022PhRvC 106b4612O doi 10 1103 PhysRevC 106 024612 S2CID 251759318 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 S2CID 4347419 Kramer K 2016 Explainer superheavy elements Chemistry World Retrieved March 15 2020 Discovery of Elements 113 and 115 Lawrence Livermore National Laboratory Archived from the original on September 11 2015 Retrieved March 15 2020 Eliav E Kaldor U Borschevsky A 2018 Electronic Structure of the Transactinide Atoms In Scott R A ed Encyclopedia of Inorganic and Bioinorganic Chemistry John Wiley amp Sons pp 1 16 doi 10 1002 9781119951438 eibc2632 ISBN 978 1 119 95143 8 S2CID 127060181 Oganessian Yu Ts Dmitriev S N Yeremin A V et al 2009 Attempt to produce the isotopes of element 108 in the fusion reaction 136Xe 136Xe Physical Review C 79 2 024608 doi 10 1103 PhysRevC 79 024608 ISSN 0556 2813 Munzenberg G Armbruster P Folger H et al 1984 The identification of element 108 PDF Zeitschrift fur Physik A 317 2 235 236 Bibcode 1984ZPhyA 317 235M doi 10 1007 BF01421260 S2CID 123288075 Archived from the original PDF on June 7 2015 Retrieved October 20 2012 Subramanian S August 28 2019 Making New Elements Doesn t Pay Just Ask This Berkeley Scientist Bloomberg Businessweek Retrieved January 18 2020 a b c d e f Ivanov D 2019 Sverhtyazhelye shagi v neizvestnoe Superheavy steps into the unknown nplus1 ru in Russian Retrieved February 2 2020 Hinde D 2017 Something new and superheavy at the periodic table The Conversation Retrieved January 30 2020 Kern B D Thompson W E Ferguson J M 1959 Cross sections for some n p and n a reactions Nuclear Physics 10 226 234 Bibcode 1959NucPh 10 226K doi 10 1016 0029 5582 59 90211 1 Wakhle A Simenel C Hinde D J et al 2015 Simenel C Gomes P R S Hinde D J et al eds Comparing Experimental and Theoretical Quasifission Mass Angle Distributions European Physical Journal Web of Conferences 86 00061 Bibcode 2015EPJWC 8600061W doi 10 1051 epjconf 20158600061 hdl 1885 148847 ISSN 2100 014X Nuclear Reactions PDF pp 7 8 Retrieved January 27 2020 Published as Loveland W D Morrissey D J Seaborg G T 2005 Nuclear Reactions Modern Nuclear Chemistry John Wiley amp Sons Inc pp 249 297 doi 10 1002 0471768626 ch10 ISBN 978 0 471 76862 3 a b Krasa A 2010 Neutron Sources for ADS Faculty of Nuclear Sciences and Physical Engineering Czech Technical University in Prague 4 8 S2CID 28796927 Wapstra A H 1991 Criteria that must be satisfied for the discovery of a new chemical element to be recognized PDF Pure and Applied Chemistry 63 6 883 doi 10 1351 pac199163060879 ISSN 1365 3075 S2CID 95737691 a b Hyde E K Hoffman D C Keller O L 1987 A History and Analysis of the Discovery of Elements 104 and 105 Radiochimica Acta 42 2 67 68 doi 10 1524 ract 1987 42 2 57 ISSN 2193 3405 S2CID 99193729 a b c d Chemistry World 2016 How to Make Superheavy Elements and Finish the Periodic Table Video Scientific American Retrieved January 27 2020 Hoffman Ghiorso amp Seaborg 2000 p 334 Hoffman Ghiorso amp Seaborg 2000 p 335 Zagrebaev Karpov amp Greiner 2013 p 3 Beiser 2003 p 432 a b Pauli N 2019 Alpha decay PDF Introductory Nuclear Atomic and Molecular Physics Nuclear Physics Part Universite libre de Bruxelles Retrieved February 16 2020 a b c d e Pauli N 2019 Nuclear fission PDF Introductory Nuclear Atomic and Molecular Physics Nuclear Physics Part Universite libre de Bruxelles Retrieved February 16 2020 Staszczak A Baran A Nazarewicz W 2013 Spontaneous fission modes and lifetimes of superheavy elements in the nuclear density functional theory Physical Review C 87 2 024320 1 arXiv 1208 1215 Bibcode 2013PhRvC 87b4320S doi 10 1103 physrevc 87 024320 ISSN 0556 2813 Audi et al 2017 pp 030001 129 030001 138 Beiser 2003 p 439 a b Beiser 2003 p 433 Audi et al 2017 p 030001 125 Aksenov N V Steinegger P Abdullin F Sh et al 2017 On the volatility of nihonium Nh Z 113 The European Physical Journal A 53 7 158 Bibcode 2017EPJA 53 158A doi 10 1140 epja i2017 12348 8 ISSN 1434 6001 S2CID 125849923 Beiser 2003 p 432 433 a b c Oganessian Yu 2012 Nuclei in the Island of Stability of Superheavy Elements Journal of Physics Conference Series 337 1 012005 1 012005 6 Bibcode 2012JPhCS 337a2005O doi 10 1088 1742 6596 337 1 012005 ISSN 1742 6596 Moller P Nix J R 1994 Fission properties of the heaviest elements PDF Dai 2 Kai Hadoron Tataikei no Simulation Symposium Tokai mura Ibaraki Japan University of North Texas Retrieved February 16 2020 a b Oganessian Yu Ts 2004 Superheavy elements Physics World 17 7 25 29 doi 10 1088 2058 7058 17 7 31 Retrieved February 16 2020 Schadel M 2015 Chemistry of the superheavy elements Philosophical Transactions of the Royal Society A Mathematical Physical and Engineering Sciences 373 2037 20140191 Bibcode 2015RSPTA 37340191S doi 10 1098 rsta 2014 0191 ISSN 1364 503X PMID 25666065 Hulet E K 1989 Biomodal spontaneous fission 50th Anniversary of Nuclear Fission Leningrad USSR Bibcode 1989nufi rept 16H Oganessian Yu Ts Rykaczewski K P 2015 A beachhead on the island of stability Physics Today 68 8 32 38 Bibcode 2015PhT 68h 32O doi 10 1063 PT 3 2880 ISSN 0031 9228 OSTI 1337838 S2CID 119531411 Grant A 2018 Weighing the heaviest elements Physics Today doi 10 1063 PT 6 1 20181113a S2CID 239775403 Howes L 2019 Exploring the superheavy elements at the end of the periodic table Chemical amp Engineering News Retrieved January 27 2020 a b Robinson A E 2019 The Transfermium Wars Scientific Brawling and Name Calling during the Cold War Distillations Retrieved February 22 2020 Populyarnaya biblioteka himicheskih elementov Siborgij ekavolfram Popular library of chemical elements Seaborgium eka tungsten n t ru in Russian Retrieved January 7 2020 Reprinted from Ekavolfram Eka tungsten Populyarnaya biblioteka himicheskih elementov Serebro Nilsborij i dalee Popular library of chemical elements Silver through nielsbohrium and beyond in Russian Nauka 1977 Nobelium Element information properties and uses Periodic Table Royal Society of Chemistry Retrieved March 1 2020 a b Kragh 2018 pp 38 39 Kragh 2018 p 40 a b Ghiorso A Seaborg G T Oganessian Yu Ts et al 1993 Responses on the report Discovery of the Transfermium elements followed by reply to the responses by Transfermium Working Group PDF Pure and Applied Chemistry 65 8 1815 1824 doi 10 1351 pac199365081815 S2CID 95069384 Archived PDF from the original on November 25 2013 Retrieved September 7 2016 Commission on Nomenclature of Inorganic Chemistry 1997 Names and symbols of transfermium elements IUPAC Recommendations 1997 PDF Pure and Applied Chemistry 69 12 2471 2474 doi 10 1351 pac199769122471 a b Hofmann S et al 1996 The new element 112 Zeitschrift fur Physik A 354 1 229 230 Bibcode 1996ZPhyA 354 229H doi 10 1007 BF02769517 S2CID 119975957 Hofmann S et al 2000 New Results on Element 111 and 112 PDF European Physical Journal A Gesellschaft fur Schwerionenforschung 14 2 147 157 Bibcode 2002EPJA 14 147H doi 10 1140 epja i2001 10119 x S2CID 8773326 Archived from the original PDF on February 27 2008 Retrieved March 2 2008 Morita K 2004 Decay of an Isotope 277112 produced by 208Pb 70Zn reaction In Penionzhkevich Yu E Cherepanov E A eds Exotic Nuclei Proceedings of the International Symposium World Scientific pp 188 191 doi 10 1142 9789812701749 0027 Sumita Takayuki Morimoto Kouji Kaji Daiya Haba Hiromitsu Ozeki Kazutaka Sakai Ryutaro Yoneda Akira Yoshida Atsushi Hasebe Hiroo Katori Kenji Sato Nozomi Wakabayashi Yasuo Mitsuoka Shin Ichi Goto Shin Ichi Murakami Masashi Kariya Yoshiki Tokanai Fuyuki Mayama Keita Takeyama Mirei Moriya Toru Ideguchi Eiji Yamaguchi Takayuki Kikunaga Hidetoshi Chiba Junsei Morita Kosuke 2013 New Result on the Production of277Cn by the208Pb 70Zn Reaction Journal of the Physical Society of Japan 82 2 024202 Bibcode 2013JPSJ 82b4202S doi 10 7566 JPSJ 82 024202 Popeko Andrey G 2016 Synthesis of superheavy elements PDF jinr ru Joint Institute for Nuclear Research Archived from the original PDF on February 4 2018 Retrieved February 4 2018 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 S2CID 97615948 Archived from the original PDF on March 9 2018 Retrieved January 9 2008 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 S2CID 95920517 Archived from the original PDF on August 22 2016 Retrieved January 9 2008 Dressler R Turler A 2001 Evidence for Isomeric States in 261Rf PDF Annual Report Paul Scherrer Institute Archived from the original PDF on July 7 2011 A New Chemical Element in the Periodic Table Gesellschaft fur Schwerionenforschung June 10 2009 Archived from the original on August 23 2009 Retrieved April 14 2012 a b c d 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 S2CID 95703833 Hofmann S Heinz S Mann R Maurer J Munzenberg 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 Schott H J Shaughnessy D A Stoyer M A Thorle 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 eds Exotic Nuclei EXON 2016 Proceedings of the International Symposium on Exotic Nuclei Exotic Nuclei pp 155 164 ISBN 9789813226555 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 New Chemical Element in the Periodic Table Science Daily June 11 2009 Element 112 shall be named copernicium Gesellschaft fur Schwerionenforschung July 14 2009 Archived from the original on July 18 2009 a b New element named copernicium BBC News July 16 2009 Retrieved February 22 2010 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 Meija Juris 2009 The need for a fresh symbol to designate copernicium Nature 461 7262 341 Bibcode 2009Natur 461 341M doi 10 1038 461341c PMID 19759598 van der Krogt P Lutetium Elementymology amp Elements Multidict Retrieved February 22 2010 Minutes Division VIII Committee meeting Glasgow 2009 PDF iupac org IUPAC 2009 Retrieved January 11 2024 Tatsumi Kazuyuki Corish John 2010 Name and symbol of the element with atomic number 112 IUPAC Recommendations 2010 PDF Pure and Applied Chemistry 82 3 753 755 doi 10 1351 PAC REC 09 08 20 Retrieved January 11 2024 IUPAC Element 112 is Named Copernicium IUPAC February 19 2010 Archived from the original on March 4 2016 Retrieved April 13 2012 Utyonkov V K Brewer N T Oganessian Yu Ts et al January 30 2018 Neutron deficient superheavy nuclei obtained in the 240Pu 48Ca reaction Physical Review C 97 14320 014320 Bibcode 2018PhRvC 97a4320U doi 10 1103 PhysRevC 97 014320 Samark Roth A Cox D M Rudolph D et al 2021 Spectroscopy along Flerovium Decay Chains Discovery of 280Ds and an Excited State in 282Cn Physical Review Letters 126 3 032503 Bibcode 2021PhRvL 126c2503S doi 10 1103 PhysRevLett 126 032503 hdl 10486 705608 PMID 33543956 S2CID 231818619 Kaji Daiya Morita Kosuke Morimoto Kouji Haba Hiromitsu Asai Masato Fujita Kunihiro Gan Zaiguo Geissel Hans Hasebe Hiroo Hofmann Sigurd Huang MingHui Komori Yukiko Ma Long Maurer Joachim Murakami Masashi Takeyama Mirei Tokanai Fuyuki Tanaka Taiki Wakabayashi Yasuo Yamaguchi Takayuki Yamaki Sayaka Yoshida Atsushi 2017 Study of the Reaction 48Ca 248Cm 296Lv at RIKEN GARIS Journal of the Physical Society of Japan 86 3 034201 1 7 Bibcode 2017JPSJ 86c4201K doi 10 7566 JPSJ 86 034201 Hofmann S Heinz S Mann R et al 2012 The reaction 48Ca 248Cm 296116 studied at the GSI SHIP The European Physical Journal A 48 5 62 Bibcode 2012EPJA 48 62H doi 10 1140 epja i2012 12062 1 S2CID 121930293 a b c Holden N E 2004 Table of the Isotopes In D R Lide ed CRC Handbook of Chemistry and Physics 85th ed CRC Press Section 11 ISBN 978 0 8493 0485 9 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 S2CID 98065999 Zagrebaev Karpov amp Greiner 2013 pp 1 15 Ninov V et al 1999 Observation of Superheavy Nuclei Produced in the Reaction of 86 Kr with 208 Pb Physical Review Letters 83 6 1104 1107 Bibcode 1999PhRvL 83 1104N doi 10 1103 PhysRevLett 83 1104 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 At Lawrence Berkeley Physicists Say a Colleague Took Them for a Ride George Johnson The New York Times 15 October 2002 Public Affairs Department October 26 2010 Six New Isotopes of the Superheavy Elements Discovered Moving Closer to Understanding the Island of Stability Berkeley Lab Retrieved April 25 2011 Subramanian S August 28 2019 Making New Elements Doesn t Pay Just Ask This Berkeley Scientist Bloomberg Businessweek Retrieved January 18 2020 a b Chemistry on the islands of stability New Scientist 11 September 1975 p 574 ISSN 1032 1233 Pitzer K S 1975 Are elements 112 114 and 118 relatively inert gases The Journal of Chemical Physics 63 2 1032 1033 doi 10 1063 1 431398 Mosyagin N S Isaev T A Titov A V 2006 Is E112 a relatively inert element Benchmark relativistic correlation study of spectroscopic constants in E112H and its cation The Journal of Chemical Physics 124 22 224302 arXiv physics 0508024 Bibcode 2006JChPh 124v4302M doi 10 1063 1 2206189 PMID 16784269 S2CID 119339584 Demissie Taye B Ruud Kenneth February 25 2017 Darmstadtium roentgenium and copernicium form strong bonds with cyanide International Journal of Quantum Chemistry 2017 e25393 doi 10 1002 qua 25393 hdl 10037 13632 Kratz Jens Volker The Impact of Superheavy Elements on the Chemical and Physical Sciences Archived June 14 2022 at the Wayback Machine 4th International Conference on the Chemistry and Physics of the Transactinide Elements 5 11 September 2011 Sochi Russia a b Eichler R Aksenov N V Belozerov A V Bozhikov G A Chepigin V I Dmitriev S N Dressler R Gaggeler H W et al 2008 Thermochemical and physical properties of element 112 Angewandte Chemie 47 17 3262 3266 doi 10 1002 anie 200705019 PMID 18338360 a b Gaston Nicola Opahle Ingo Gaggeler Heinz W Schwerdtfeger Peter 2007 Is eka mercury element 112 a group 12 metal Angewandte Chemie 46 10 1663 1666 doi 10 1002 anie 200604262 PMID 17397075 Retrieved November 5 2013 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 Bibcode 2018SSCom 269 16G doi 10 1016 j ssc 2017 10 009 Retrieved March 28 2018 Cencarikova Hana Legut Dominik 2018 The effect of relativity on stability of Copernicium phases their electronic structure and mechanical properties Physica B 536 576 582 arXiv 1810 01955 Bibcode 2018PhyB 536 576C doi 10 1016 j physb 2017 11 035 S2CID 119100368 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 a b c d e Gaggeler H W 2007 Gas Phase Chemistry of Superheavy Elements PDF Paul Scherrer Institute pp 26 28 Archived from the original PDF on February 20 2012 Moody Ken 2013 Synthesis of Superheavy Elements In Schadel Matthias Shaughnessy Dawn eds The Chemistry of Superheavy Elements 2nd ed Springer Science amp Business Media pp 24 28 ISBN 9783642374661 Zaitsevskii A van Wullen 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 February 17 2018 Annual Report 2015 Laboratory of Radiochemistry and Environmental Chemistry PDF Paul Scherrer Institute 2015 p 3 Bibliography editAudi G Kondev F G Wang M Huang W J Naimi S 2017 The NUBASE2016 evaluation of nuclear properties Chinese Physics C 41 3 030001 Bibcode 2017ChPhC 41c0001A doi 10 1088 1674 1137 41 3 030001 Beiser A 2003 Concepts of modern physics 6th ed McGraw Hill ISBN 978 0 07 244848 1 OCLC 48965418 Hoffman D C Ghiorso A Seaborg G T 2000 The Transuranium People The Inside Story World Scientific ISBN 978 1 78 326244 1 Kragh H 2018 From Transuranic to Superheavy Elements A Story of Dispute and Creation Springer ISBN 978 3 319 75813 8 Zagrebaev Valeriy Karpov Alexander Greiner Walter 2013 Future of superheavy element research Which nuclei could be synthesized within the next few years PDF 11th International Conference on Nucleus Nucleus Collisions NN2012 Journal of Physics Conference Series Vol 420 IOP Publishing doi 10 1088 1742 6596 420 1 012001 Retrieved August 20 2013 External links edit nbsp Look up copernicium in Wiktionary the free dictionary nbsp Wikimedia Commons has media related to copernicium Copernicium at The Periodic Table of Videos University of Nottingham Retrieved from https en wikipedia org w index php title Copernicium amp oldid 1205397818, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.