Tennessine

From Wikipedia, the free encyclopedia

Tennessine, 117Ts
Tennessine
Pronunciation/ˈtɛnəsn/[1] (TEN-ə-seen)
Appearancesemimetallic (predicted)[2]
Mass number[294]
Tennessine 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
At

Ts

(Usu)
livermoriumtennessineoganesson
Atomic number (Z)117
Groupgroup 17 (halogens)
Periodperiod 7
Block  p-block
Electron configuration[Rn] 5f14 6d10 7s2 7p5 (predicted)[3]
Electrons per shell2, 8, 18, 32, 32, 18, 7 (predicted)
Physical properties
Phase at STPsolid (predicted)[3][4]
Melting point623–823 K ​(350–550 °C, ​662–1022 °F) (predicted)[3]
Boiling point883 K ​(610 °C, ​1130 °F) (predicted)[3]
Density (near r.t.)7.1–7.3 g/cm3 (extrapolated)[4]
Atomic properties
Oxidation states(−1), (+1), (+3), (+5) (predicted)[2][3]
Ionization energies
  • 1st: 742.9 kJ/mol (predicted)[5]
  • 2nd: 1435.4 kJ/mol (predicted)[5]
  • 3rd: 2161.9 kJ/mol (predicted)[5]
  • (more)
Atomic radiusempirical: 138 pm (predicted)[4]
Covalent radius156–157 pm (extrapolated)[4]
Other properties
Natural occurrencesynthetic
CAS Number54101-14-3
History
Namingafter Tennessee region
DiscoveryJoint Institute for Nuclear Research, Lawrence Livermore National Laboratory, Vanderbilt University and Oak Ridge National Laboratory (2009)
Isotopes of tennessine
Main isotopes[6] Decay
abun­dance half-life (t1/2) mode pro­duct
293Ts synth 25 ms[6][7] α 289Mc
294Ts synth 51 ms[8] α 290Mc
 Category: Tennessine
| references

Tennessine is a synthetic chemical element; it has symbol Ts and atomic number 117. It has the second-highest atomic number and joint-highest atomic mass of all known elements, and is the penultimate element of the 7th period of the periodic table.

The discovery of tennessine was officially announced in Dubna, Russia, by a Russian–American collaboration in April 2010, which makes it the most recently discovered element as of 2024. One of its daughter isotopes was created directly in 2011, partially confirming the results of the experiment. The experiment itself was repeated successfully by the same collaboration in 2012 and by a joint German–American team in May 2014. In December 2015, the Joint Working Party of the International Union of Pure and Applied Chemistry (IUPAC) and the International Union of Pure and Applied Physics (IUPAP), which evaluates claims of discovery of new elements, recognized the element and assigned the priority to the Russian–American team. In June 2016, the IUPAC published a declaration stating that the discoverers had suggested the name tennessine after Tennessee, United States, a name which was officially adopted in November 2016.[a]

Tennessine may be located in the "island of stability", a concept that explains why some superheavy elements are more stable compared to an overall trend of decreasing stability for elements beyond bismuth on the periodic table. The synthesized tennessine atoms have lasted tens and hundreds of milliseconds. In the periodic table, tennessine is expected to be a member of group 17, the halogens.[b] Some of its properties may differ significantly from those of the lighter halogens due to relativistic effects. As a result, tennessine is expected to be a volatile metal that neither forms anions nor achieves high oxidation states. A few key properties, such as its melting and boiling points and its first ionization energy, are nevertheless expected to follow the periodic trends of the halogens.

Introduction[edit]

Synthesis of superheavy nuclei[edit]

A graphic depiction of a nuclear fusion reaction
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[c] atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size[d] into one; roughly, the more unequal the two nuclei in terms of mass, the greater the possibility that the two react.[15] 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.[16] 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.[16]

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.[16][17] This happens because during the attempted formation of a single nucleus, electrostatic repulsion tears apart the nucleus that is being formed.[16] 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.[e] 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.[16]

External videos
video icon Visualization of unsuccessful nuclear fusion, based on calculations from the Australian National University[19]

The resulting merger is an excited state[20]—termed a compound nucleus—and thus it is very unstable.[16] To reach a more stable state, the temporary merger may fission without formation of a more stable nucleus.[21] 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.[21] 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.[22][f]

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.[24] In the separator, the newly produced nucleus is separated from other nuclides (that of the original beam and any other reaction products)[g] 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.[24] The transfer takes about 10−6 seconds; in order to be detected, the nucleus must survive this long.[27] The nucleus is recorded again once its decay is registered, and the location, the energy, and the time of the decay are measured.[24]

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.[28] 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.[29][30] Superheavy nuclei are thus theoretically predicted[31] and have so far been observed[32] to predominantly decay via decay modes that are caused by such repulsion: alpha decay and spontaneous fission.[h] Almost all alpha emitters have over 210 nucleons,[34] and the lightest nuclide primarily undergoing spontaneous fission has 238.[35] In both decay modes, nuclei are inhibited from decaying by corresponding energy barriers for each mode, but they can be tunnelled through.[29][30]

Apparatus for creation of superheavy elements
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.[36]

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.[37] Spontaneous fission is caused by electrostatic repulsion tearing the nucleus apart and produces various nuclei in different instances of identical nuclei fissioning.[30] 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),[38] and by 30 orders of magnitude from thorium (element 90) to fermium (element 100).[39] 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.[30][40] 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.[30][40] 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.[41] Experiments on lighter superheavy nuclei,[42] as well as those closer to the expected island,[38] have shown greater than previously anticipated stability against spontaneous fission, showing the importance of shell effects on nuclei.[i]

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.[j] (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.)[24] 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).[k] Spontaneous fission, however, produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[l]

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.[m]

History[edit]

Pre-discovery[edit]

In December 2004, the Joint Institute for Nuclear Research (JINR) team in Dubna, Moscow Oblast, Russia, proposed a joint experiment with the Oak Ridge National Laboratory (ORNL) in Oak Ridge, Tennessee, United States, to synthesize element 117 — so called for the 117 protons in its nucleus. Their proposal involved fusing a berkelium (element 97) target and a calcium (element 20) beam, conducted via bombardment of the berkelium target with calcium nuclei:[53] this would complete a set of experiments done at the JINR on the fusion of actinide targets with a calcium-48 beam, which had thus far produced the new elements 113116 and 118. The ORNL—then the world's only producer of berkelium—could not then provide the element, as they had temporarily ceased production,[53] and re-initiating it would be too costly.[54] Plans to synthesize element 117 were suspended in favor of the confirmation of element 118, which had been produced earlier in 2002 by bombarding a californium target with calcium.[55] The required berkelium-249 is a by-product in californium-252 production, and obtaining the required amount of berkelium was an even more difficult task than obtaining that of californium, as well as costly: It would cost around 3.5 million dollars, and the parties agreed to wait for a commercial order of californium production, from which berkelium could be extracted.[54][56]

The JINR team sought to use berkelium because calcium-48, the isotope of calcium used in the beam, has 20 protons and 28 neutrons, making a neutron–proton ratio of 1.4; and it is the lightest stable or near-stable nucleus with such a large neutron excess. Thanks to the neutron excess, the resulting nuclei were expected to be heavier and closer to the sought-after island of stability.[n] Of the aimed for 117 protons, calcium has 20, and thus they needed to use berkelium, which has 97 protons in its nucleus.[6]

In February 2005, the leader of the JINR team — Yuri Oganessian — presented a colloquium at ORNL. Also in attendance were representatives of Lawrence Livermore National Laboratory, who had previously worked with JINR on the discovery of elements 113–116 and 118, and Joseph Hamilton of Vanderbilt University, a collaborator of Oganessian.[58]

Hamilton checked if the ORNL high-flux reactor produced californium for a commercial order: The required berkelium could be obtained as a by-product. He learned that it did not and there was no expectation for such an order in the immediate future. Hamilton kept monitoring the situation, making the checks once in a while. (Later, Oganessian referred to Hamilton as "the father of 117" for doing this work.)[58]

Discovery[edit]

ORNL resumed californium production in spring 2008. Hamilton noted the restart during the summer and made a deal on subsequent extraction of berkelium[59] (the price was about $600,000).[60] During a September 2008 symposium at Vanderbilt University in Nashville, Tennessee, celebrating his 50th year on the Physics faculty, Hamilton introduced Oganessian to James Roberto (then the deputy director for science and technology at ORNL).[61] They established a collaboration among JINR, ORNL, and Vanderbilt.[56] Clarice Phelps was part of ORNL's team that collaborated with JINR;[62] this is particularly notable as because of it the IUPAC recognizes her as the first African-American woman to be involved with the discovery of a chemical element.[62][63][64][65] The eventual collaborating institutions also included The University of Tennessee (Knoxville), Lawrence Livermore National Laboratory, The Research Institute for Advanced Reactors (Russia), and The University of Nevada (Las Vegas).[66]

A very small sample of a blue liquid in a plastic pipette held by a hand wearing heavy protection equipment
The berkelium target used for the synthesis (in solution)

In November 2008, the U.S. Department of Energy, which had oversight over the reactor in Oak Ridge, allowed the scientific use of the extracted berkelium.[67]

The production lasted 250 days and ended in late December 2008,[68] resulting in 22 milligrams of berkelium, enough to perform the experiment.[69] In January 2009, the berkelium was removed from ORNL's High Flux Isotope Reactor;[67] it was subsequently cooled for 90 days and then processed at ORNL's Radiochemical Engineering and Development Center to separate and purify the berkelium material, which took another 90 days.[56] Its half-life is only 330 days: this means, after that time, half the berkelium produced would have decayed. Because of this, the berkelium target had to be quickly transported to Russia; for the experiment to be viable, it had to be completed within six months of its departure from the United States.[56] The target was packed into five lead containers to be flown from New York to Moscow.[56] Russian customs officials twice refused to let the target enter the country because of missing or incomplete paperwork. Over the span of a few days, the target traveled over the Atlantic Ocean five times.[56] On its arrival in Russia in June 2009, the berkelium was immediately transferred to Research Institute of Atomic Reactors (RIAR) in Dimitrovgrad, Ulyanovsk Oblast, where it was deposited as a 300-nanometer-thin layer on a titanium film.[68] In July 2009, it was transported to Dubna,[68] where it was installed in the particle accelerator at the JINR.[69] The calcium-48 beam was generated by chemically extracting the small quantities of calcium-48 present in naturally occurring calcium, enriching it 500 times.[citation needed] This work was done in the closed town of Lesnoy, Sverdlovsk Oblast, Russia.[67]

The experiment began in late July 2009.[67] In January 2010, scientists at the Flerov Laboratory of Nuclear Reactions announced internally that they had detected the decay of a new element with atomic number 117 via two decay chains: one of an odd–odd isotope undergoing 6 alpha decays before spontaneous fission, and one of an odd–even isotope undergoing 3 alpha decays before fission.[70] The obtained data from the experiment was sent to the LLNL for further analysis.[71] On 9 April 2010, an official report was released in the journal Physical Review Letters identifying the isotopes as 294117 and 293117, which were shown to have half-lives on the order of tens or hundreds of milliseconds. The work was signed by all parties involved in the experiment to some extent: JINR, ORNL, LLNL, RIAR, Vanderbilt, the University of Tennessee (Knoxville, Tennessee, U.S.), and the University of Nevada (Las Vegas, Nevada, U.S.), which provided data analysis support.[72] The isotopes were formed as follows:[73][o]

249
97
Bk
+ 48
20
Ca
297117* → 294117 + 3 1
0

n
(1 event)
249
97
Bk
+ 48
20
Ca
297117* → 293117 + 4 1
0

n
(5 events)

Confirmation[edit]

Decay chain of the atoms produced in the original experiment. The figures near the arrows describe experimental (black) and theoretical (blue) values for the lifetime and energy of each decay. Lifetimes may be converted to half-lives by multiplying by ln 2.[73]

All daughter isotopes (decay products) of element 117 were previously unknown;[73] therefore, their properties could not be used to confirm the claim of discovery. In 2011, when one of the decay products (289115) was synthesized directly, its properties matched those measured in the claimed indirect synthesis from the decay of element 117.[74] The discoverers did not submit a claim for their findings in 2007–2011 when the Joint Working Party was reviewing claims of discoveries of new elements.[75]

The Dubna team repeated the experiment in 2012, creating seven atoms of element 117 and confirming their earlier synthesis of element 118 (produced after some time when a significant quantity of the berkelium-249 target had beta decayed to californium-249). The results of the experiment matched the previous outcome;[8] the scientists then filed an application to register the element.[citation needed] In May 2014, a joint German–American collaboration of scientists from the ORNL and the GSI Helmholtz Center for Heavy Ion Research in Darmstadt, Hessen, Germany, claimed to have confirmed discovery of the element.[7][76] The team repeated the Dubna experiment using the Darmstadt accelerator, creating two atoms of element 117.[7]

In December 2015, the JWP officially recognized the discovery of 293117 on account of the confirmation of the properties of its daughter 289115,[77] and thus the listed discoverers — JINR, LLNL, and ORNL — were given the right to suggest an official name for the element. (Vanderbilt was left off the initial list of discoverers in an error that was later corrected.)[78]

In May 2016, Lund University (Lund, Scania, Sweden) and GSI cast some doubt on the syntheses of elements 115 and 117. The decay chains assigned to 289115, the isotope instrumental in the confirmation of the syntheses of elements 115 and 117, were found based on a new statistical method to be too different to belong to the same nuclide with a reasonably high probability. The reported 293117 decay chains approved as such by the JWP were found to require splitting into individual data sets assigned to different isotopes of element 117. It was also found that the claimed link between the decay chains reported as from 293117 and 289115 probably did not exist. (On the other hand, the chains from the non-approved isotope 294117 were found to be congruent.) The multiplicity of states found when nuclides that are not even–even undergo alpha decay is not unexpected and contributes to the lack of clarity in the cross-reactions. This study criticized the JWP report for overlooking subtleties associated with this issue, and considered it "problematic" that the only argument for the acceptance of the discoveries of elements 115 and 117 was a link they considered to be doubtful.[79][80]

On 8 June 2017, two members of the Dubna team published a journal article answering these criticisms, analysing their data on the nuclides 293117 and 289115 with widely accepted statistical methods, noted that the 2016 studies indicating non-congruence produced problematic results when applied to radioactive decay: they excluded from the 90% confidence interval both average and extreme decay times, and the decay chains that would be excluded from the 90% confidence interval they chose were more probable to be observed than those that would be included. The 2017 reanalysis concluded that the observed decay chains of 293117 and 289115 were consistent with the assumption that only one nuclide was present at each step of the chain, although it would be desirable to be able to directly measure the mass number of the originating nucleus of each chain as well as the excitation function of the 243Am + 48Ca reaction.[81]

Naming[edit]

Main campus of Hamilton's workplace, Vanderbilt University, one of the institutions named as co-discoverers of tennessine

Using Mendeleev's nomenclature for unnamed and undiscovered elements, element 117 should be known as eka-astatine. Using the 1979 recommendations by the International Union of Pure and Applied Chemistry (IUPAC), the element was temporarily called ununseptium (symbol Uus) until its discovery was confirmed and a permanent name chosen; the temporary name was formed from Latin roots "one", "one", and "seven", a reference to the element's atomic number 117.[82] Many scientists in the field called it "element 117", with the symbol E117, (117), or 117.[3] According to guidelines of IUPAC valid at the moment of the discovery approval, the permanent names of new elements should have ended in "-ium"; this included element 117, even if the element was a halogen, which traditionally have names ending in "-ine";[83] however, the new recommendations published in 2016 recommended using the "-ine" ending for all new group 17 elements.[84]

After the original synthesis in 2010, Dawn Shaughnessy of LLNL and Oganessian declared that naming was a sensitive question, and it was avoided as far as possible.[85] However, Hamilton declared that year, "I was crucial in getting the group together and in getting the 249Bk target essential for the discovery. As a result of that, I'm going to get to name the element. I can't tell you the name, but it will bring distinction to the region."[72] (Hamilton teaches at Vanderbilt University in Nashville, Tennessee, U.S.) In a 2015 interview, Oganessian, after telling the story of the experiment, said, "and the Americans named this a tour de force, they had demonstrated they could do [this] with no margin for error. Well, soon they will name the 117th element."[86]

In March 2016, the discovery team agreed on a conference call involving representatives from the parties involved on the name "tennessine" for element 117.[58] In June 2016, IUPAC published a declaration stating the discoverers had submitted their suggestions for naming the new elements 115, 117, and 118 to the IUPAC; the suggestion for the element 117 was tennessine, with a symbol of Ts, after "the region of Tennessee".[a] The suggested names were recommended for acceptance by the IUPAC Inorganic Chemistry Division; formal acceptance was set to occur after a five-month term following publishing of the declaration expires.[87] In November 2016, the names, including tennessine, were formally accepted. Concerns that the proposed symbol Ts may clash with a notation for the tosyl group used in organic chemistry were rejected, following existing symbols bearing such dual meanings: Ac (actinium and acetyl) and Pr (praseodymium and propyl).[88] The naming ceremony for moscovium, tennessine, and oganesson was held on 2 March 2017 at the Russian Academy of Sciences in Moscow; a separate ceremony for tennessine alone had been held at ORNL in January 2017.[89]

Predicted properties[edit]

Other than nuclear properties, no properties of tennessine or its compounds have been measured; this is due to its extremely limited and expensive production[60] and the fact that it decays very quickly. Properties of tennessine remain unknown and only predictions are available.

Nuclear stability and isotopes[edit]

The stability of nuclei quickly decreases with the increase in atomic number after curium, element 96, whose half-life is four orders of magnitude longer than that of any subsequent element. All isotopes with an atomic number above 101 undergo radioactive decay with half-lives of less than 30 hours. No elements with atomic numbers above 82 (after lead) have stable isotopes.[90] This is because of the ever-increasing Coulomb repulsion of protons, so that the strong nuclear force cannot hold the nucleus together against spontaneous fission for long. Calculations suggest that in the absence of other stabilizing factors, elements with more than 104 protons should not exist.[91] However, researchers in the 1960s suggested that the closed nuclear shells around 114 protons and 184 neutrons should counteract this instability, creating an "island of stability" where nuclides could have half-lives reaching thousands or millions of years. While scientists have still not reached the island, the mere existence of the superheavy elements (including tennessine) confirms that this stabilizing effect is real, and in general the known superheavy nuclides become exponentially longer-lived as they approach the predicted location of the island.[92][93] Tennessine is the second-heaviest element created so far, and all its known isotopes have half-lives of less than one second. Nevertheless, this is longer than the values predicted prior to their discovery: the predicted lifetimes for 293Ts and 294Ts used in the discovery paper were 10 ms and 45 ms respectively, while the observed lifetimes were 21 ms and 112 ms respectively.[73] The Dubna team believes that the synthesis of the element is direct experimental proof of the existence of the island of stability.[94]

A 2D graph with rectangular cells colored in black-and-white colors, spanning from the llc to the urc, with cells mostly becoming lighter closer to the latter
A chart of nuclide stability as used by the Dubna team in 2010. Characterized isotopes are shown with borders. According to the discoverers, the synthesis of element 117 serves as definite proof of the existence of the "island of stability" (circled).[94]

It has been calculated that the isotope 295Ts would have a half-life of about 18 milliseconds, and it may be possible to produce this isotope via the same berkelium–calcium reaction used in the discoveries of the known isotopes, 293Ts and 294Ts. The chance of this reaction producing 295Ts is estimated to be, at most, one-seventh the chance of producing 294Ts.[27][95][96] Calculations using a quantum tunneling model predict the existence of several isotopes of tennessine up to 303Ts. The most stable of these is expected to be 296Ts with an alpha-decay half-life of 40 milliseconds.[97] A liquid drop model study on the element's isotopes shows similar results; it suggests a general trend of increasing stability for isotopes heavier than 301Ts, with partial half-lives exceeding the age of the universe for the heaviest isotopes like 335Ts when beta decay is not considered.[98] Lighter isotopes of tennessine may be produced in the 243Am+50Ti reaction, which was considered as a contingency plan by the Dubna team in 2008 if 249Bk proved unavailable.[99]

Atomic and physical[edit]

Tennessine is expected to be a member of group 17 in the periodic table, below the five halogens; fluorine, chlorine, bromine, iodine, and astatine, each of which has seven valence electrons with a configuration of ns2np5.[100][p] For tennessine, being in the seventh period (row) of the periodic table, continuing the trend would predict a valence electron configuration of 7s27p5,[3] and it would therefore be expected to behave similarly to the halogens in many respects that relate to this electronic state. However, going down group 17, the metallicity of the elements increases; for example, iodine already exhibits a metallic luster in the solid state, and astatine is expected to be a metal.[101] As such, an extrapolation based on periodic trends would predict tennessine to be a rather volatile metal.[102]

Black-on-transparent graph, width greater than height, with the main part of the graph being filled with short horizontal stripes
Atomic energy levels of outermost s, p, and d electrons of chlorine (d orbitals not applicable), bromine, iodine, astatine, and tennessine

Calculations have confirmed the accuracy of this simple extrapolation, although experimental verification of this is currently impossible as the half-lives of the known tennessine isotopes are too short.[102] Significant differences between tennessine and the previous halogens are likely to arise, largely due to spin–orbit interaction—the mutual interaction between the motion and spin of electrons. The spin–orbit interaction is especially strong for the superheavy elements because their electrons move faster—at velocities comparable to the speed of light—than those in lighter atoms.[103] In tennessine atoms, this lowers the 7s and the 7p electron energy levels, stabilizing the corresponding electrons, although two of the 7p electron energy levels are more stabilized than the other four.[104] The stabilization of the 7s electrons is called the inert pair effect; the effect that separates the 7p subshell into the more-stabilized and the less-stabilized parts is called subshell splitting. Computational chemists understand the split as a change of the second (azimuthal) quantum number l from 1 to 1/2 and 3/2 for the more-stabilized and less-stabilized parts of the 7p subshell, respectively.[105][q] For many theoretical purposes, the valence electron configuration may be represented to reflect the 7p subshell split as 7s2
7p2
1/2
7p3
3/2
.[3]

Differences for other electron levels also exist. For example, the 6d electron levels (also split in two, with four being 6d3/2 and six being 6d5/2) are both raised, so they are close in energy to the 7s ones,[104] although no 6d electron chemistry has ever been predicted for tennessine. The difference between the 7p1/2 and 7p3/2 levels is abnormally high; 9.8 eV.[104] Astatine's 6p subshell split is only 3.8 eV,[104] and its 6p1/2 chemistry has already been called "limited".[106] These effects cause tennessine's chemistry to differ from those of its upper neighbors (see below).

Tennessine's first ionization energy—the energy required to remove an electron from a neutral atom—is predicted to be 7.7 eV, lower than those of the halogens, again following the trend.[3] Like its neighbors in the periodic table, tennessine is expected to have the lowest electron affinity—energy released when an electron is added to the atom—in its group; 2.6 or 1.8 eV.[3] The electron of the hypothetical hydrogen-like tennessine atom—oxidized so it has only one electron, Ts116+—is predicted to move so quickly that its mass is 1.90 times that of a non-moving electron, a feature attributable to relativistic effects. For comparison, the figure for hydrogen-like astatine is 1.27 and the figure for hydrogen-like iodine is 1.08.[107] Simple extrapolations of relativity laws indicate a contraction of atomic radius.[107] Advanced calculations show that the radius of an tennessine atom that has formed one covalent bond would be 165 pm, while that of astatine would be 147 pm.[108] With the seven outermost electrons removed, tennessine is finally smaller; 57 pm[3] for tennessine and 61 pm[109] for astatine.

The melting and boiling points of tennessine are not known; earlier papers predicted about 350–500 °C and 550 °C, respectively,[3] or 350–550 °C and 610 °C, respectively.[110] These values exceed those of astatine and the lighter halogens, following periodic trends. A later paper predicts the boiling point of tennessine to be 345 °C[111] (that of astatine is estimated as 309 °C,[112] 337 °C,[113] or 370 °C,[114] although experimental values of 230 °C[115] and 411 °C[109] have been reported). The density of tennessine is expected to be between 7.1 and 7.3 g/cm3.[4]

Chemical[edit]

Skeletal model of a planar molecule with a central atom (iodine) symmetrically bonded to three (fluorine) atoms to form a big right-angled T
IF
3
has a T-shape configuration.
Skeletal model of a trigonal molecule with a central atom (tennessine) symmetrically bonded to three peripheral (fluorine) atoms
TsF
3
is predicted to have a trigonal configuration.

The known isotopes of tennessine, 293Ts and 294Ts, are too short-lived to allow for chemical experimentation at present. Nevertheless, many chemical properties of tennessine have been calculated.[116] Unlike the lighter group 17 elements, tennessine may not exhibit the chemical behavior common to the halogens.[9] For example, fluorine, chlorine, bromine, and iodine routinely accept an electron to achieve the more stable electronic configuration of a noble gas, obtaining eight electrons (octet) in their valence shells instead of seven.[117] This ability weakens as atomic weight increases going down the group; tennessine would be the least willing group 17 element to accept an electron. Of the oxidation states it is predicted to form, −1 is expected to be the least common.[3] The standard reduction potential of the Ts/Ts couple is predicted to be −0.25 V; this value is negative, unlike for all the lighter halogens.[2]

There is another opportunity for tennessine to complete its octet—by forming a covalent bond. Like the halogens, when two tennessine atoms meet they are expected to form a Ts–Ts bond to give a diatomic molecule. Such molecules are commonly bound via single sigma bonds between the atoms; these are different from pi bonds, which are divided into two parts, each shifted in a direction perpendicular to the line between the atoms, and opposite one another rather than being located directly between the atoms they bind. Sigma bonding has been calculated to show a great antibonding character in the At2 molecule and is not as favorable energetically. Tennessine is predicted to continue the trend; a strong pi character should be seen in the bonding of Ts2.[3][118] The molecule tennessine chloride (TsCl) is predicted to go further, being bonded with a single pi bond.[118]

Aside from the unstable −1 state, three more oxidation states are predicted; +5, +3, and +1. The +1 state should be especially stable because of the destabilization of the three outermost 7p3/2 electrons, forming a stable, half-filled subshell configuration;[3] astatine shows similar effects.[119] The +3 state should be important, again due to the destabilized 7p3/2 electrons.[110] The +5 state is predicted to be uncommon because the 7p1/2 electrons are oppositely stabilized.[3] The +7 state has not been shown—even computationally—to be achievable. Because the 7s electrons are greatly stabilized, it has been hypothesized that tennessine effectively has only five valence electrons.[120]

The simplest possible tennessine compound would be the monohydride, TsH. The bonding is expected to be provided by a 7p3/2 electron of tennessine and the 1s electron of hydrogen. The non-bonding nature of the 7p1/2 spinor is because tennessine is expected not to form purely sigma or pi bonds.[121] Therefore, the destabilized (thus expanded) 7p3/2 spinor is responsible for bonding.[122] This effect lengthens the TsH molecule by 17 picometers compared with the overall length of 195 pm.[121] Since the tennessine p electron bonds are two-thirds sigma, the bond is only two-thirds as strong as it would be if tennessine featured no spin–orbit interactions.[121] The molecule thus follows the trend for halogen hydrides, showing an increase in bond length and a decrease in dissociation energy compared to AtH.[3] The molecules TlTs and NhTs may be viewed analogously, taking into account an opposite effect shown by the fact that the element's p1/2 electrons are stabilized. These two characteristics result in a relatively small dipole moment (product of difference between electric charges of atoms and displacement of the atoms) for TlTs; only 1.67 D,[r] the positive value implying that the negative charge is on the tennessine atom. For NhTs, the strength of the effects are predicted to cause a transfer of the electron from the tennessine atom to the nihonium atom, with the dipole moment value being −1.80 D.[124] The spin–orbit interaction increases the dissociation energy of the TsF molecule because it lowers the electronegativity of tennessine, causing the bond with the extremely electronegative fluorine atom to have a more ionic character.[121] Tennessine monofluoride should feature the strongest bonding of all group 17 monofluorides.[121]

VSEPR theory predicts a bent-T-shaped molecular geometry for the group 17 trifluorides. All known halogen trifluorides have this molecular geometry and have a structure of AX3E2—a central atom, denoted A, surrounded by three ligands, X, and two unshared electron pairs, E. If relativistic effects are ignored, TsF3 should follow its lighter congeners in having a bent-T-shaped molecular geometry. More sophisticated predictions show that this molecular geometry would not be energetically favored for TsF3, predicting instead a trigonal planar molecular geometry (AX3E0). This shows that VSEPR theory may not be consistent for the superheavy elements.[120] The TsF3 molecule is predicted to be significantly stabilized by spin–orbit interactions; a possible rationale may be the large difference in electronegativity between tennessine and fluorine, giving the bond a partially ionic character.[120]

Notes[edit]

  1. ^ a b The declaration by the IUPAC mentioned "the contribution of the Tennessee region (emphasis added), including Oak Ridge National Laboratory, Vanderbilt University, and the University of Tennessee at Knoxville, Tennessee, to superheavy element research, including the production and chemical separation of unique actinide target materials for superheavy element synthesis at ORNL’s High Flux Isotope Reactor (HFIR) and Radiochemical Engineering Development Center (REDC)".
  2. ^ The term "group 17" refers to a column in the periodic table starting with fluorine. The term "halogen" is sometimes considered as synonymous, but sometimes it instead relates to a common set of chemical and physical properties shared by fluorine, chlorine, bromine, iodine, and astatine, all of which precede tennessine in group 17. Unlike the other group 17 members, tennessine might not be a halogen under this stricter definition.[9]
  3. ^ 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[10] or 112;[11] sometimes, the term is presented an equivalent to the term "transactinide", which puts an upper limit before the beginning of the hypothetical superactinide series).[12] 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.
  4. ^ 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.[13] 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.[14]
  5. ^ 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.[18]
  6. ^ This figure also marks the generally accepted upper limit for lifetime of a compound nucleus.[23]
  7. ^ 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.[25] 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.[26]
  8. ^ Not all decay modes are caused by electrostatic repulsion. For example, beta decay is caused by the weak interaction.[33]
  9. ^ 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.[38]
  10. ^ 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.[43] The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[44] 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).[45]
  11. ^ 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).[34] 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.
  12. ^ Spontaneous fission was discovered by Soviet physicist Georgy Flerov,[46] a leading scientist at JINR, and thus it was a "hobbyhorse" for the facility.[47] 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.[23] They thus preferred to link new isotopes to the already known ones by successive alpha decays.[46]
  13. ^ For instance, element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm, Stockholm County, Sweden.[48] 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.[49] 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.[49] JINR insisted that they were the first to create the element and suggested a name of their own for the new element, joliotium;[50] the Soviet name was also not accepted (JINR later referred to the naming of the element 102 as "hasty").[51] This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements, signed 29 September 1992.[51] The name "nobelium" remained unchanged on account of its widespread usage.[52]
  14. ^ Although stable isotopes of the lightest elements usually have a neutron–proton ratio close or equal to one (for example, the only stable isotope of aluminium has 13 protons and 14 neutrons,[6] making a neutron–proton ratio of 1.077), stable isotopes of heavier elements have higher neutron–proton ratios, increasing with the number of protons. For example, iodine's only stable isotope has 53 protons and 74 neutrons, giving neutron–proton ratio of 1.396, gold's only stable isotope has 79 protons and 118 neutrons, yielding a neutron–proton ratio of 1.494, and plutonium's most stable isotope has 94 protons and 150 neutrons, and a neutron–proton ratio of 1.596.[6] This trend[57] is expected to make it difficult to synthesize the most stable isotopes of super-heavy elements as the neutron–proton ratios of the elements they are synthesized from will be too low.
  15. ^ A nuclide is commonly denoted by the chemical element's symbol immediately preceded by the mass number as a superscript and the atomic number as a subscript. Neutrons are represented as nuclides with atomic mass 1, atomic number 0, and symbol n. Outside the context of nuclear equations, the atomic number is sometimes omitted. An asterisk denotes an extremely short-lived (or even non-existent) intermediate stage of the reaction.
  16. ^ The letter n stands for the number of the period (horizontal row in the periodic table) the element belongs to. The letters "s" and "p" denote the s and p atomic orbitals, and the subsequent superscript numbers denote the numbers of electrons in each. Hence the notation ns2np5 means that the valence shells of lighter group 17 elements are composed of two s electrons and five p electrons, all located in the outermost electron energy level.
  17. ^ The quantum number corresponds to the letter in the electron orbital name: 0 to s, 1 to p, 2 to d, etc. See azimuthal quantum number for more information.
  18. ^ For comparison, the values for the ClF, HCl, SO, HF, and HI molecules are 0.89 D, 1.11 D, 1.55 D, 1.83 D, and 1.95 D. Values for molecules which do not form at standard conditions, namely GeSe, SnS, TlF, BaO, and NaCl, are 1.65 D, ~3.2 D, 4.23 D, 7.95 D, and 9.00 D.[123]

References[edit]

  1. ^ Ritter, Malcolm (9 June 2016). "Periodic table elements named for Moscow, Japan, Tennessee". Associated Press. Retrieved 19 December 2017.
  2. ^ a b c Fricke, Burkhard (1975). "Superheavy elements: a prediction of their chemical and physical properties". Recent Impact of Physics on Inorganic Chemistry. Structure and Bonding. 21: 89–144. doi:10.1007/BFb0116498. ISBN 978-3-540-07109-9. Retrieved 4 October 2013.
  3. ^ a b c d e f g h i j k l m n o p q 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.
  4. ^ a b c d e Bonchev, D.; Kamenska, V. (1981). "Predicting the Properties of the 113–120 Transactinide Elements". Journal of Physical Chemistry. 85 (9): 1177–1186. doi:10.1021/j150609a021.
  5. ^ a b c Chang, Zhiwei; Li, Jiguang; Dong, Chenzhong (2010). "Ionization Potentials, Electron Affinities, Resonance Excitation Energies, Oscillator Strengths, And Ionic Radii of Element Uus (Z = 117) and Astatine". J. Phys. Chem. A. 2010 (114): 13388–94. Bibcode:2010JPCA..11413388C. doi:10.1021/jp107411s.
  6. ^ a b c d e 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.
  7. ^ a b c Khuyagbaatar, J.; Yakushev, A.; Düllmann, Ch. E.; et al. (2014). "48Ca+249Bk Fusion Reaction Leading to Element Z=117: Long-Lived α-Decaying 270Db and Discovery of 266Lr". Physical Review Letters. 112 (17): 172501. Bibcode:2014PhRvL.112q2501K. doi:10.1103/PhysRevLett.112.172501. PMID 24836239.
  8. ^ a b Oganessian, Yu. Ts.; et al. (2013). "Experimental studies of the 249Bk + 48Ca reaction including decay properties and excitation function for isotopes of element 117, and discovery of the new isotope 277Mt". Physical Review C. 87 (5): 054621. Bibcode:2013PhRvC..87e4621O. doi:10.1103/PhysRevC.87.054621.
  9. ^ a b "Superheavy Element 117 Confirmed – On the Way to the "Island of Stability"". GSI Helmholtz Centre for Heavy Ion Research. Archived from the original on 3 August 2018. Retrieved 26 July 2015.
  10. ^ Krämer, K. (2016). "Explainer: superheavy elements". Chemistry World. Retrieved 15 March 2020.
  11. ^ "Discovery of Elements 113 and 115". Lawrence Livermore National Laboratory. Archived from the original on 11 September 2015. Retrieved 15 March 2020.
  12. ^ 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.
  13. ^ 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.
  14. ^ Münzenberg, G.; Armbruster, P.; Folger, H.; et al. (1984). "The identification of element 108" (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 7 June 2015. Retrieved 20 October 2012.
  15. ^ Subramanian, S. (28 August 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 18 January 2020.
  16. ^ a b c d e f Ivanov, D. (2019). "Сверхтяжелые шаги в неизвестное" [Superheavy steps into the unknown]. nplus1.ru (in Russian). Retrieved 2 February 2020.
  17. ^ Hinde, D. (2017). "Something new and superheavy at the periodic table". The Conversation. Retrieved 30 January 2020.
  18. ^ 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.
  19. ^ 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.
  20. ^ "Nuclear Reactions" (PDF). pp. 7–8. Retrieved 27 January 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.
  21. ^ 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.
  22. ^ 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.
  23. ^ 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.
  24. ^ a b c d Chemistry World (2016). "How to Make Superheavy Elements and Finish the Periodic Table [Video]". Scientific American. Retrieved 27 January 2020.
  25. ^ Hoffman, Ghiorso & Seaborg 2000, p. 334.
  26. ^ Hoffman, Ghiorso & Seaborg 2000, p. 335.
  27. ^ a b Zagrebaev, Karpov & Greiner 2013, p. 3.
  28. ^ Beiser 2003, p. 432.
  29. ^ a b Pauli, N. (2019). "Alpha decay" (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part). Université libre de Bruxelles. Retrieved 16 February 2020.
  30. ^ 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 16 February 2020.
  31. ^ 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.
  32. ^ Audi et al. 2017, pp. 030001-129–030001-138.
  33. ^ Beiser 2003, p. 439.
  34. ^ a b Beiser 2003, p. 433.
  35. ^ Audi et al. 2017, p. 030001-125.
  36. ^ 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.
  37. ^ Beiser 2003, p. 432–433.
  38. ^ 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.
  39. ^ 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 16 February 2020.
  40. ^ a b Oganessian, Yu. Ts. (2004). "Superheavy elements". Physics World. 17 (7): 25–29. doi:10.1088/2058-7058/17/7/31. Retrieved 16 February 2020.
  41. ^ 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.
  42. ^ Hulet, E. K. (1989). Biomodal spontaneous fission. 50th Anniversary of Nuclear Fission, Leningrad, USSR. Bibcode:1989nufi.rept...16H.
  43. ^ 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.
  44. ^ Grant, A. (2018). "Weighing the heaviest elements". Physics Today. doi:10.1063/PT.6.1.20181113a. S2CID 239775403.
  45. ^ Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table". Chemical & Engineering News. Retrieved 27 January 2020.
  46. ^ a b Robinson, A. E. (2019). "The Transfermium Wars: Scientific Brawling and Name-Calling during the Cold War". Distillations. Retrieved 22 February 2020.
  47. ^ "Популярная библиотека химических элементов. Сиборгий (экавольфрам)" [Popular library of chemical elements. Seaborgium (eka-tungsten)]. n-t.ru (in Russian). Retrieved 7 January 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.
  48. ^ "Nobelium - Element information, properties and uses | Periodic Table". Royal Society of Chemistry. Retrieved 1 March 2020.
  49. ^ a b Kragh 2018, pp. 38–39.
  50. ^ Kragh 2018, p. 40.
  51. ^ 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 25 November 2013. Retrieved 7 September 2016.
  52. ^ 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.
  53. ^ a b Cabage, B. (2010). "International team discovers element 117" (Press release). Oak Ridge National Laboratory. Archived from the original on 23 September 2015. Retrieved 26 June 2017.
  54. ^ a b "Vanderbilt physicist plays pivotal role in discovery of new super-heavy element" (Press release). Vanderbilt University. April 2010. Retrieved 12 June 2016.
  55. ^ Oganessian, Yu.Ts.; Utyonkov, V.K.; Lobanov, Yu.V.; Abdullin, F.Sh.; Polyakov, A.N.; Shirokovsky, I.V.; et al. (2002). "Results from the first 249Cf+48Ca experiment" (PDF). JINR Communication. Retrieved 23 September 2015.
  56. ^ a b c d e f Bardi, J. S. (2010). "An Atom at the End of the Material World". Inside Science. Retrieved 3 January 2015.
  57. ^ Karpov, A. V.; Zagrebaev, V. I.; Palenzuela, Y. Martinez; Greiner, Walter (2013). "Superheavy Nuclei: Decay and Stability". Exciting Interdisciplinary Physics. FIAS Interdisciplinary Science Series. p. 69. doi:10.1007/978-3-319-00047-3_6. ISBN 978-3-319-00046-6.
  58. ^ a b c "What it takes to make a new element". Chemistry World. Retrieved 3 December 2016.
  59. ^ Witze, Alexandra (2010). "The backstory behind a new element". Science News. Retrieved 12 June 2016.
  60. ^ a b Subramanian, S. (28 August 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 8 March 2024.
  61. ^ Siner, Emily (2016). "How scientists plan to enshrine Tennessee on the periodic table of elements". National Public Radio. Retrieved 7 March 2017.
  62. ^ a b "Clarice Phelps". IUPAC 100.
  63. ^ "PT of Younger Chemists". IUPAC 100.
  64. ^ https://web.archive.org/web/20190729202324/https://www.oakridger.com/news/20190729/two-ornl-researchers-featured-on-periodic-table-of-younger-chemists
  65. ^ https://pubs.aip.org/physicstoday/online/31578
  66. ^ "The Discovery of Tennessine" (PDF). Oak Ridge National Laboratory. Retrieved 11 June 2023.
  67. ^ a b c d Roberto, James (2010). "The discovery of element 117" (PDF) (Press release). Oak Ridge National Laboratory. Archived from the original (PDF) on 21 October 2016. Retrieved 26 June 2017.
  68. ^ a b c "For the Press" (Press release). Joint Institute for Nuclear Research. 2010. Archived from the original on 4 March 2016. Retrieved 28 July 2015.
  69. ^ a b Stark, A.M. (2010). "International team discovers element 117" (Press release). DOE / Lawrence Livermore National Laboratory. Retrieved 29 November 2012.
  70. ^ Greiner, W. (2010). Recommendations (PDF). 31st meeting, PAC for nuclear physics. p. 6. Archived from the original (PDF) on 14 April 2010.
  71. ^ "Nations work together to discover new element". DOE Office of Science. U.S. Department of Energy (Press release). U.S. Department of Energy. 2011. Retrieved 5 January 2016.
  72. ^ a b "Heaviest in the world". Arts and Science Magazine. Vanderbilt University. November 2011. Archived from the original on 3 May 2016. Retrieved 12 June 2016.
  73. ^ a b c d Oganessian, Yu.Ts.; Abdullin, F.Sh.; Bailey, P.D.; Benker, D.E.; Bennett, M.E.; Dmitriev, S.N.; et al. (2010). "Synthesis of a new element with atomic number Z = 117". Physical Review Letters. 104 (14): 142502. Bibcode:2010PhRvL.104n2502O. doi:10.1103/PhysRevLett.104.142502. PMID 20481935. S2CID 3263480.
  74. ^ Molchanov, E. (2011). В лабораториях ОИЯИ. Возвращение к дубнию [In JINR labs. Returning to dubnium] (in Russian). JINR. Retrieved 9 November 2011.
  75. ^ Barber, R.C.; Karol, P.J.; Nakahara, H.; Vardaci, E.; Vogt, E.W. (2011). "Discovery of the elements with atomic numbers greater than or equal to 113". Pure and Applied Chemistry. IUPAC Technical Report. 83 (7): 1485–1498. doi:10.1351/PAC-REP-10-05-01. S2CID 98065999.
  76. ^ Chow, D. (1 May 2014). "New super-heavy element 117 confirmed by scientists". Live Science. Retrieved 2 May 2014.
  77. ^ "Discovery and assignment of elements with atomic numbers 113, 115, 117 and 118" (Press release). IUPAC. 2015. Archived from the original on 7 February 2016. Retrieved 4 January 2016.
  78. ^ Karol, Paul J.; Barber, Robert C.; Sherrill, Bradley M.; Vardaci, Emanuele; Yamazaki, Toshimitsu (22 December 2015). "Discovery of the elements with atomic numbers Z = 113, 115, and 117" (PDF). Pure Appl. Chem. IUPAC Technical Report. 88 (1–2): 139–153. doi:10.1515/pac-2015-0502. S2CID 101634372. Retrieved 2 April 2016.
  79. ^ Forsberg, U.; Rudolph, D.; Fahlander, C.; Golubev, P.; Sarmiento, L.G.; Åberg, S.; Block, M.; Düllmann, Ch.E.; Heßberger, F.P.; Kratz, J.V.; Yakushev, A. (9 July 2016). "A new assessment of the alleged link between element 115 and element 117 decay chains" (PDF). Physics Letters B. 760 (2016): 293–296. Bibcode:2016PhLB..760..293F. doi:10.1016/j.physletb.2016.07.008. Retrieved 2 April 2016.
  80. ^ Forsberg, Ulrika; Fahlander, Claes; Rudolph, Dirk (2016). Congruence of decay chains of elements 113, 115, and 117 (PDF). Nobel Symposium NS160 – Chemistry and Physics of Heavy and Superheavy Elements. doi:10.1051/epjconf/201613102003.
  81. ^ Zlokazov, V.B.; Utyonkov, V.K. (8 June 2017). "Analysis of decay chains of superheavy nuclei produced in the 249Bk + 48Ca and 243Am + 48Ca reactions". Journal of Physics G: Nuclear and Particle Physics. 44 (7): 075107. Bibcode:2017JPhG...44g5107Z. doi:10.1088/1361-6471/aa7293.
  82. ^ Chatt, J. (1979). "Recommendations for the naming of elements of atomic numbers greater than 100". Pure Appl. Chem. 51 (2): 381–384. doi:10.1351/pac197951020381.
  83. ^ Koppenol, W.H. (2002). "Naming of new elements" (PDF). Pure and Applied Chemistry. IUPAC Recommendations 2002. 74 (5): 787–791. doi:10.1351/pac200274050787. S2CID 95859397.
  84. ^ Koppenol, Willem H.; Corish, John; García-Martínez, Javier; Meija, Juris; Reedijk, Jan (2016). "How to name new chemical elements" (PDF). Pure and Applied Chemistry. IUPAC Recommendations 2016. 88 (4): 401–405. doi:10.1515/pac-2015-0802. hdl:10045/55935. S2CID 102245448.
  85. ^ Glanz, J. (2010). "Scientists discover heavy new element". Department of Chemistry (Press release). Oregon State University. Retrieved 5 January 2016.
  86. ^ Oganessian, Yu.Ts. (10 October 2015). "Гамбургский счет" [Hamburg reckoning] (Interview) (in Russian). Interviewed by Orlova, O. Public Television of Russia. Archived from the original on 11 November 2021. Retrieved 18 January 2020.
  87. ^ "IUPAC Is Naming The Four New Elements Nihonium, Moscovium, Tennessine, and Oganesson" (Press release). IUPAC. 8 June 2016. Retrieved 8 June 2016.
  88. ^ "IUPAC Announces the Names of the Elements 113, 115, 117, and 118 - IUPAC | International Union of Pure and Applied Chemistry". IUPAC | International Union of Pure and Applied Chemistry. 30 November 2016. Retrieved 30 November 2016.
  89. ^ Fedorova, Vera (3 March 2017). "At the inauguration ceremony of the new elements of the periodic table of D.I. Mendeleev". jinr.ru. Joint Institute for Nuclear Research. Retrieved 4 February 2018.
  90. ^ de Marcillac, P.; Coron, N.; Dambier, G.; et al. (2003). "Experimental detection of α-particles from the radioactive decay of natural bismuth". Nature. 422 (6934): 876–878. Bibcode:2003Natur.422..876D. doi:10.1038/nature01541. PMID 12712201. S2CID 4415582.
  91. ^ Möller, P. (2016). "The limits of the nuclear chart set by fission and alpha decay" (PDF). EPJ Web of Conferences. 131: 03002:1–8. Bibcode:2016EPJWC.13103002M. doi:10.1051/epjconf/201613103002.
  92. ^ Considine, G.D.; Kulik, Peter H. (2002). Van Nostrand's scientific encyclopedia (9th ed.). Wiley-Interscience. ISBN 978-0-471-33230-5. OCLC 223349096.
  93. ^ Oganessian, Yu. Ts.; Sobiczewski, A.; Ter-Akopian, G. M. (9 January 2017). "Superheavy nuclei: from predictions to discovery". Physica Scripta. 92 (2): 023003–1–21. Bibcode:2017PhyS...92b3003O. doi:10.1088/1402-4896/aa53c1. S2CID 125713877.
  94. ^ a b "Element 117 is synthesized". JINR. 2010. Retrieved 28 June 2015.
  95. ^ Zhao-Qing, F.; Gen-Ming, Jin; Ming-Hui, Huang; et al. (2007). "Possible Way to Synthesize Superheavy Element Z = 117". Chinese Physics Letters. 24 (9): 2551. arXiv:0708.0159. Bibcode:2007ChPhL..24.2551F. doi:10.1088/0256-307X/24/9/024. S2CID 8778306.
  96. ^ Zhao-Qing, F.; Jina, Gen-Ming; Li, Jun-Qing; et al. (2009). "Production of heavy and superheavy nuclei in massive fusion reactions". Nuclear Physics A. 816 (1–4): 33. arXiv:0803.1117. Bibcode:2009NuPhA.816...33F. doi:10.1016/j.nuclphysa.2008.11.003. S2CID 18647291.
  97. ^ Chowdhury, R. P.; Samanta, C.; Basu, D. N. (2008). "Search for long lived heaviest nuclei beyond the valley of stability". Physical Review C. 77 (4): 044603. arXiv:0802.3837. Bibcode:2008PhRvC..77d4603C. doi:10.1103/PhysRevC.77.044603. S2CID 119207807.
  98. ^ Duarte, S. B.; Tavares, O. A. P.; Gonçalves, M.; et al. (September 2004). "Half-life prediction for decay modes for superheavy nuclei" (PDF). Journal of Physics G: Nuclear and Particle Physics. Notas de Física. Centro Brasileiro de Pesquisas Físicas. 30 (CBPF-NF-022/04): 1487–1494. Bibcode:2004JPhG...30.1487D. doi:10.1088/0954-3899/30/10/014. ISSN 0029-3865.
  99. ^ Utyonkov, V. K. (12 February 2008). "Синтез новых элементов 113-118 в реакциях полного слияния 48Ca + 238U-249Cf" [Synthesis of new elements 113–118 in complete fusion reactions 48Ca + 238U–249Cf] (PDF). nuclphys.sinp.msu.ru. Retrieved 28 April 2017.
  100. ^ Dhingra, A. (1 December 1999). The Sterling Dictionary Of Chemistry. Sterling Publishers Pvt. Ltd. p. 187. ISBN 978-81-7359-123-5. Retrieved 23 July 2015.
  101. ^ Hermann, A.; Hoffmann, R.; Ashcroft, N. W. (2013). "Condensed Astatine: Monatomic and Metallic". Physical Review Letters. 111 (11): 116404-1–116404-5. Bibcode:2013PhRvL.111k6404H. doi:10.1103/PhysRevLett.111.116404. PMID 24074111.
  102. ^ a b GSI (14 December 2015). "Research Program – Highlights". superheavies.de. GSI. Archived from the original on 13 May 2020. Retrieved 9 November 2016. If this trend were followed, element 117 would likely be a rather volatile metal. Fully relativistic calculations agree with this expectation, however, they are in need of experimental confirmation.
  103. ^ Thayer 2010, pp. 63–64.
  104. ^ a b c d Fægri Jr., K.; Saue, T. (2001). "Diatomic molecules between very heavy elements of group 13 and group 17: A study of relativistic effects on bonding". The Journal of Chemical Physics. 115 (6): 2456. Bibcode:2001JChPh.115.2456F. doi:10.1063/1.1385366.
  105. ^ Thayer 2010, pp. 63–67.
  106. ^ Thayer 2010, p. 79.
  107. ^ a b Thayer 2010, p. 64.
  108. ^ Pyykkö, P.; Atsumi, M. (22 December 2008). "Molecular Single-Bond Covalent Radii for Elements 1-118". Chemistry: A European Journal. 15 (1): 186–197. doi:10.1002/chem.200800987. PMID 19058281.
  109. ^ a b Sharma, B. K. (2001). Nuclear and radiation chemistry (7th ed.). Krishna Prakashan Media. p. 147. ISBN 978-81-85842-63-9. Retrieved 9 November 2012.
  110. ^ a b Seaborg, Glenn T. (1994). Modern alchemy. World Scientific. p. 172. ISBN 978-981-02-1440-1.
  111. ^ Takahashi, N. (2002). "Boiling points of the superheavy elements 117 and 118". Journal of Radioanalytical and Nuclear Chemistry. 251 (2): 299–301. doi:10.1023/A:1014880730282. S2CID 93096903.
  112. ^ Luig, H.; Keller, C.; Wolf, W.; et al. (2005). "Radionuclides". In Ullmann, F. (ed.). Encyclopedia of industrial chemistry. Wiley-VCH. p. 23. doi:10.1002/14356007.a22_499. ISBN 978-3-527-30673-2.
  113. ^ Punter, J.; Johnson, R.; Langfield, S. (2006). The essentials of GCSE OCR Additional science for specification B. Letts and Lonsdale. p. 36. ISBN 978-1-905129-73-7.
  114. ^ Wiberg, E.; Wiberg, N.; Holleman, A. F. (2001). Inorganic chemistry. Academic Press. p. 423. ISBN 978-0-12-352651-9.
  115. ^ Otozai, K.; Takahashi, N. (1982). "Estimation of the chemical form and the boiling point of elementary astatine by radiogas-chromatography". Radiochimica Acta. 31 (3‒4): 201‒203. doi:10.1524/ract.1982.31.34.201. S2CID 100363889.
  116. ^ Moody, Ken (30 November 2013). "Synthesis of Superheavy Elements". In Schädel, Matthias; Shaughnessy, Dawn (eds.). The Chemistry of Superheavy Elements (2nd ed.). Springer Science & Business Media. pp. 24–8. ISBN 9783642374661.
  117. ^ Bader, R. F. W. "An introduction to the electronic structure of atoms and molecules". McMaster University. Archived from the original on 12 October 2007. Retrieved 18 January 2008.
  118. ^ a b Pershina 2010, p. 504.
  119. ^ Thayer 2010, p. 84.
  120. ^ a b c Bae, Ch.; Han, Y.-K.; Lee, Yo. S. (18 January 2003). "Spin−Orbit and Relativistic Effects on Structures and Stabilities of Group 17 Fluorides EF3 (E = I, At, and Element 117): Relativity Induced Stability for the D3h Structure of (117)F3". The Journal of Physical Chemistry A. 107 (6): 852–858. Bibcode:2003JPCA..107..852B. doi:10.1021/jp026531m.
  121. ^ a b c d e Han, Y.-K.; Bae, Cheolbeom; Son, Sang-Kil; et al. (2000). "Spin-orbit effects on the transactinide p-block element monohydrides MH (M=element 113-118)". Journal of Chemical Physics. 112 (6): 2684–2691. Bibcode:2000JChPh.112.2684H. doi:10.1063/1.480842. S2CID 9959620.
  122. ^ Stysziński 2010, pp. 144–146.
  123. ^ Lide, D. R. (2003). "Section 9, Molecular Structure and Spectroscopy". CRC Handbook of Chemistry and Physics (84th ed.). CRC Press. pp. 9–45, 9–46. ISBN 978-0-8493-0484-2.
  124. ^ Stysziński 2010, pp. 139–146.

Bibliography[edit]