Oganesson

Source: Wikipedia, the free encyclopedia.

Oganesson, 118Og
Oganesson
Pronunciation
Appearancemetallic (predicted)
Mass number[294]
Oganesson 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
Rn

Og

(Usb)
tennessineoganessonununennium
Discovery
Joint Institute for Nuclear Research and Lawrence Livermore National Laboratory (2002)
Isotopes of oganesson
Main isotopes[12] Decay
abun­dance half-life (t1/2) mode pro­duct
294Og synth 0.7 ms[13][14] α
290Lv
SF
 Category: Oganesson
| references

Oganesson is a synthetic chemical element; it has symbol Og and atomic number 118. It was first synthesized in 2002 at the Joint Institute for Nuclear Research (JINR) in Dubna, near Moscow, Russia, by a joint team of Russian and American scientists. In December 2015, it was recognized as one of four new elements by the Joint Working Party of the international scientific bodies IUPAC and IUPAP. It was formally named on 28 November 2016.[15][16] The name honors the nuclear physicist Yuri Oganessian, who played a leading role in the discovery of the heaviest elements in the periodic table. It is one of only two elements named after a person who was alive at the time of naming, the other being seaborgium, and the only element whose eponym is alive as of 2024.[17][a]

Oganesson has the highest atomic number and highest

noble gases
).

Introduction

Synthesis of superheavy nuclei

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[b] atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size[c] into one; roughly, the more unequal the two nuclei in terms of mass, the greater the possibility that the two react.[25] 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.[26] 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.[26]

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.[26][27] This happens because during the attempted formation of a single nucleus, electrostatic repulsion tears apart the nucleus that is being formed.[26] 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.[d] 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.[26]

External videos
video icon Visualization of unsuccessful nuclear fusion, based on calculations from the Australian National University[29]

The resulting merger is an excited state[30]—termed a compound nucleus—and thus it is very unstable.[26] To reach a more stable state, the temporary merger may fission without formation of a more stable nucleus.[31] 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.[31] 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.[32][e]

Decay and detection

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.[34] In the separator, the newly produced nucleus is separated from other nuclides (that of the original beam and any other reaction products)[f] 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.[34] The transfer takes about 10−6 seconds; in order to be detected, the nucleus must survive this long.[37] The nucleus is recorded again once its decay is registered, and the location, the energy, and the time of the decay are measured.[34]

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.[38] 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.[39][40] Superheavy nuclei are thus theoretically predicted[41] and have so far been observed[42] to predominantly decay via decay modes that are caused by such repulsion: alpha decay and spontaneous fission.[g] Almost all alpha emitters have over 210 nucleons,[44] and the lightest nuclide primarily undergoing spontaneous fission has 238.[45] In both decay modes, nuclei are inhibited from decaying by corresponding energy barriers for each mode, but they can be tunnelled through.[39][40]

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.[46]

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.

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.[40][50] 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.[40][50] 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.[51] Experiments on lighter superheavy nuclei,[52] as well as those closer to the expected island,[48] have shown greater than previously anticipated stability against spontaneous fission, showing the importance of shell effects on nuclei.[h]

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.[i] (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.)[34] 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).[j] Spontaneous fission, however, produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[k]

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.[l]

History

Early speculation

The possibility of a seventh noble gas, after helium, neon, argon, krypton, xenon, and radon, was considered almost as soon as the noble gas group was discovered. Danish chemist Hans Peter Jørgen Julius Thomsen predicted in April 1895, the year after the discovery of argon, that there was a whole series of chemically inert gases similar to argon that would bridge the halogen and alkali metal groups: he expected that the seventh of this series would end a 32-element period which contained thorium and uranium and have an atomic weight of 292, close to the 294 now known for the first and only confirmed isotope of oganesson.[63] Danish physicist Niels Bohr noted in 1922 that this seventh noble gas should have atomic number 118 and predicted its electronic structure as 2, 8, 18, 32, 32, 18, 8, matching modern predictions.[64] Following this, German chemist Aristid von Grosse wrote an article in 1965 predicting the likely properties of element 118.[11] It was 107 years from Thomsen's prediction before oganesson was successfully synthesized, although its chemical properties have not been investigated to determine if it behaves as the heavier congener of radon.[65] In a 1975 article, American chemist Kenneth Pitzer suggested that element 118 should be a gas or volatile liquid due to relativistic effects.[66]

Unconfirmed discovery claims

In late 1998, Polish physicist Robert Smolańczuk published calculations on the fusion of atomic nuclei towards the synthesis of superheavy atoms, including oganesson.[67] His calculations suggested that it might be possible to make element 118 by fusing lead with krypton under carefully controlled conditions, and that the fusion probability (cross section) of that reaction would be close to the lead–chromium reaction that had produced element 106, seaborgium. This contradicted predictions that the cross sections for reactions with lead or bismuth targets would go down exponentially as the atomic number of the resulting elements increased.[67]

In 1999, researchers at Lawrence Berkeley National Laboratory made use of these predictions and announced the discovery of elements 118 and 116, in a paper published in Physical Review Letters,[68] and very soon after the results were reported in Science.[69] The researchers reported that they had performed the reaction

86
36
Kr
293
118
Og
+
n
.

In 2001, they published a retraction after researchers at other laboratories were unable to duplicate the results and the Berkeley lab could not duplicate them either.[70] In June 2002, the director of the lab announced that the original claim of the discovery of these two elements had been based on data fabricated by principal author Victor Ninov.[71][72] Newer experimental results and theoretical predictions have confirmed the exponential decrease in cross sections with lead and bismuth targets as the atomic number of the resulting nuclide increases.[73]

Discovery reports

daughter isotope. The fraction of atoms undergoing spontaneous fission
(SF) is given in green.

The first genuine decay of atoms of oganesson was observed in 2002 at the Joint Institute for Nuclear Research (JINR) in Dubna, Russia, by a joint team of Russian and American scientists. Headed by Yuri Oganessian, a Russian nuclear physicist of Armenian ethnicity, the team included American scientists from the Lawrence Livermore National Laboratory in California.[74] The discovery was not announced immediately, because the decay energy of 294Og matched that of 212mPo, a common impurity produced in fusion reactions aimed at producing superheavy elements, and thus announcement was delayed until after a 2005 confirmatory experiment aimed at producing more oganesson atoms.[75] The 2005 experiment used a different beam energy (251 MeV instead of 245 MeV) and target thickness (0.34 mg/cm2 instead of 0.23 mg/cm2).[13] On 9 October 2006, the researchers announced[13] that they had indirectly detected a total of three (possibly four) nuclei of oganesson-294 (one or two in 2002[76] and two more in 2005) produced via collisions of californium-249 atoms and calcium-48 ions.[77][78][79][80][81]

294
118
Og
+ 3
n
.

In 2011, IUPAC evaluated the 2006 results of the Dubna–Livermore collaboration and concluded: "The three events reported for the Z = 118 isotope have very good internal redundancy but with no anchor to known nuclei do not satisfy the criteria for discovery".[82]

Because of the very small

false positive, since the chance that the detections were random events was estimated to be less than one part in 100000.[84]

In the experiments, the alpha-decay of three atoms of oganesson was observed. A fourth decay by direct spontaneous fission was also proposed. A half-life of 0.89 ms was calculated: 294
Og
decays into 290
Lv
by alpha decay. Since there were only three nuclei, the half-life derived from observed lifetimes has a large uncertainty: 0.89+1.07
−0.31
 ms
.[13]

294
118
Og
290
116
Lv
+ 4
2
He

The identification of the 294
Og
nuclei was verified by separately creating the putative

245
Cm with 48
Ca
ions,

245
96
Cm
+ 48
20
Ca
290
116
Lv
+ 3

n
,

and checking that the 290
Lv
decay matched the

282
Cn, which will undergo spontaneous fission.[13]

Confirmation

In December 2015, the Joint Working Party of international scientific bodies International Union of Pure and Applied Chemistry (IUPAC) and International Union of Pure and Applied Physics (IUPAP) recognized the element's discovery and assigned the priority of the discovery to the Dubna–Livermore collaboration.[85] This was on account of two 2009 and 2010 confirmations of the properties of the granddaughter of 294Og, 286Fl, at the Lawrence Berkeley National Laboratory, as well as the observation of another consistent decay chain of 294Og by the Dubna group in 2012. The goal of that experiment had been the synthesis of 294Ts via the reaction 249Bk(48Ca,3n), but the short half-life of 249Bk resulted in a significant quantity of the target having decayed to 249Cf, resulting in the synthesis of oganesson instead of tennessine.[86]

From 1 October 2015 to 6 April 2016, the Dubna team performed a similar experiment with 48Ca projectiles aimed at a mixed-isotope californium target containing 249Cf, 250Cf, and 251Cf, with the aim of producing the heavier oganesson isotopes 295Og and 296Og. Two beam energies at 252 MeV and 258 MeV were used. Only one atom was seen at the lower beam energy, whose decay chain fitted the previously known one of 294Og (terminating with spontaneous fission of 286Fl), and none were seen at the higher beam energy. The experiment was then halted, as the glue from the sector frames covered the target and blocked evaporation residues from escaping to the detectors.[87] The production of 293Og and its daughter 289Lv, as well as the even heavier isotope 297Og, is also possible using this reaction. The isotopes 295Og and 296Og may also be produced in the fusion of 248Cm with 50Ti projectiles.[87][88][89] A search beginning in summer 2016 at RIKEN for 295Og in the 3n channel of this reaction was unsuccessful, though the study is planned to resume; a detailed analysis and cross section limit were not provided. These heavier and likely more stable isotopes may be useful in probing the chemistry of oganesson.[90][91]

Naming

Element 118 was named after Yuri Oganessian, a pioneer in the discovery of synthetic elements, with the name oganesson (Og). Oganessian and the decay chain of oganesson-294 were pictured on a stamp of Armenia issued on 28 December 2017.

Using Mendeleev's nomenclature for unnamed and undiscovered elements, oganesson is sometimes known as eka-radon (until the 1960s as eka-emanation, emanation being the old name for radon).[11] In 1979, IUPAC assigned the systematic placeholder name ununoctium to the undiscovered element, with the corresponding symbol of Uuo,[92] and recommended that it be used until after confirmed discovery of the element.[93] 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 called it "element 118", with the symbol of E118, (118), or even simply 118.[4]

Before the retraction in 2001, the researchers from Berkeley had intended to name the element ghiorsium (Gh), after Albert Ghiorso (a leading member of the research team).[94]

The Russian discoverers reported their synthesis in 2006. According to IUPAC recommendations, the discoverers of a new element have the right to suggest a name.

Flyorov Laboratory of Nuclear Reactions at JINR was the only facility in the world which could achieve this result.[97] These names were later suggested for element 114 (flerovium) and element 116 (moscovium).[98] Flerovium became the name of element 114; the final name proposed for element 116 was instead livermorium,[99] with moscovium later being proposed and accepted for element 115 instead.[17]

Traditionally, the names of all

group 18 elements, regardless of whether they turn out to have the chemical properties of a noble gas.[101]

The scientists involved in the discovery of element 118, as well as those of 117 and 115, held a conference call on 23 March 2016 to decide their names. Element 118 was the last to be decided upon; after Oganessian was asked to leave the call, the remaining scientists unanimously decided to have the element "oganesson" after him. Oganessian was a pioneer in superheavy element research for sixty years reaching back to the field's foundation: his team and his proposed techniques had led directly to the synthesis of elements 107 through 118. Mark Stoyer, a nuclear chemist at the LLNL, later recalled, "We had intended to propose that name from Livermore, and things kind of got proposed at the same time from multiple places. I don't know if we can claim that we actually proposed the name, but we had intended it."[102]

In internal discussions, IUPAC asked the JINR if they wanted the element to be spelled "oganeson" to match the Russian spelling more closely. Oganessian and the JINR refused this offer, citing the Soviet-era practice of transliterating names into the Latin alphabet under the rules of the French language ("Oganessian" is such a transliteration) and arguing that "oganesson" would be easier to link to the person.[103][m] In June 2016, IUPAC announced that the discoverers planned to give the element the name oganesson (symbol: Og). The name became official on 28 November 2016.[17] In 2017, Oganessian commented on the naming:[104]

For me, it is an honour. The discovery of element 118 was by scientists at the Joint Institute for Nuclear Research in Russia and at the Lawrence Livermore National Laboratory in the US, and it was my colleagues who proposed the name oganesson. My children and grandchildren have been living in the US for decades, but my daughter wrote to me to say that she did not sleep the night she heard because she was crying.[104]

— Yuri Oganessian

The naming ceremony for moscovium, tennessine, and oganesson was held on 2 March 2017 at the Russian Academy of Sciences in Moscow.[105]

In a 2019 interview, when asked what it was like to see his name in the periodic table next to

Mendeleev, the Curies, and Rutherford, Oganessian responded:[103]

Not like much! You see, not like much. It is customary in science to name something new after its discoverer. It's just that there are few elements, and this happens rarely. But look at how many equations and theorems in mathematics are named after somebody. And in medicine? Alzheimer, Parkinson. There's nothing special about it.

Characteristics

Other than nuclear properties, no properties of oganesson or its compounds have been measured; this is due to its extremely limited and expensive production[106] and the fact that it decays very quickly. Thus only predictions are available.

Nuclear stability and isotopes

Oganesson (row 118) is slightly above the "island of stability" (white ellipse) and thus its nuclei are slightly more stable than otherwise predicted.

The stability of nuclei quickly decreases with the increase in atomic number after

radioactive, decaying via alpha decay and spontaneous fission,[111][112] with a half-life that appears to be less than a millisecond. Nonetheless, this is still longer than some predicted values.[113][114]

Calculations using a quantum-tunneling model predict the existence of several heavier isotopes of oganesson with alpha-decay half-lives close to 1 ms.[115][116]

Theoretical calculations done on the synthetic pathways for, and the half-life of, other

stable than the synthesized isotope 294Og, most likely 293Og, 295Og, 296Og, 297Og, 298Og, 300Og and 302Og (the last reaching the N = 184 shell closure).[113][117] Of these, 297Og might provide the best chances for obtaining longer-lived nuclei,[113][117] and thus might become the focus of future work with this element. Some isotopes with many more neutrons, such as some located around 313Og, could also provide longer-lived nuclei.[118]

In a

quantum-tunneling model, the alpha decay half-life of 294
Og
was predicted to be 0.66+0.23
−0.18
 ms
[113] with the experimental Q-value published in 2004.[119] Calculation with theoretical Q-values from the macroscopic-microscopic model of Muntian–Hofman–Patyk–Sobiczewski gives somewhat lower but comparable results.[120]

Calculated atomic and physical properties

Oganesson is a member of

closed outer valence shell in which its valence electrons are arranged in a 7s27p6 configuration.[3]

Consequently, some expect oganesson to have similar physical and chemical properties to other members of its group, most closely resembling the noble gas above it in the periodic table, radon.[122] Following the

anion Og by 9%, thus confirming the importance of these corrections in superheavy elements.[123] 2022 calculations expect the electron affinity of oganesson to be 0.080(6) eV.[8]

standard conditions, though still with a rather low melting point.[5][19]

Oganesson is expected to have an extremely broad polarizability, almost double that of radon.[3] Because of its tremendous polarizability, oganesson is expected to have an anomalously low first ionization energy of about 860 kJ/mol, similar to that of cadmium and less than those of iridium, platinum, and gold. This is significantly smaller than the values predicted for darmstadtium, roentgenium, and copernicium, although it is greater than that predicted for flerovium.[128] Its second ionization energy should be around 1560 kJ/mol.[8] Even the shell structure in the nucleus and electron cloud of oganesson is strongly impacted by relativistic effects: the valence and core electron subshells in oganesson are expected to be "smeared out" in a homogeneous Fermi gas of electrons, unlike those of the "less relativistic" radon and xenon (although there is some incipient delocalisation in radon), due to the very strong spin–orbit splitting of the 7p orbital in oganesson.[129] A similar effect for nucleons, particularly neutrons, is incipient in the closed-neutron-shell nucleus 302Og and is strongly in force at the hypothetical superheavy closed-shell nucleus 472164, with 164 protons and 308 neutrons.[129] Studies have also predicted that due to increasing electrostatic forces, oganesson may have a semibubble structure in proton density, having few protons at the center of its nucleus.[130][131] Moreover, spin–orbit effects may cause bulk oganesson to be a semiconductor, with a band gap of 1.5±0.6 eV predicted. All the lighter noble gases are insulators instead: for example, the band gap of bulk radon is expected to be 7.1±0.5 eV.[132]

Predicted compounds

Skeletal model of a planar molecule with a central atom symmetrically bonded to four peripheral (fluorine) atoms.
XeF
4
has a square planar molecular geometry.
Skeletal model of a terahedral molecule with a central atom (oganesson) symmetrically bonded to four peripheral (fluorine) atoms.
OgF
4
is predicted to have a tetrahedral molecular geometry.

The only confirmed isotope of oganesson, 294Og, has much too short a half-life to be chemically investigated experimentally. Therefore, no compounds of oganesson have been synthesized yet.

oxidize and therefore, the most common oxidation state would be 0 (as for the noble gases);[133] nevertheless, this appears not to be the case.[65]

Calculations on the

dissociation energy of 6 kJ/mol, roughly 4 times of that of Rn
2
.[3] Most strikingly, it was calculated to have a bond length shorter than in Rn
2
by 0.16 Å, which would be indicative of a significant bonding interaction.[3] On the other hand, the compound OgH+ exhibits a dissociation energy (in other words proton affinity of oganesson) that is smaller than that of RnH+.[3]

The bonding between oganesson and

electropositivity.[138] Oganesson is predicted to be sufficiently electropositive[138] to form an Og–Cl bond with chlorine.[7]

A compound of oganesson and tennessine, OgTs4, has been predicted to be potentially stable chemically.[139]

See also

Notes

  1. ^ The names einsteinium and fermium for elements 99 and 100 were proposed when their namesakes (Albert Einstein and Enrico Fermi respectively) were still alive, but were not made official until Einstein and Fermi had died.[18]
  2. superactinide series).[22]
    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.
  3. pb.[23] 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.[24]
  4. ^ 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.[28]
  5. ^ This figure also marks the generally accepted upper limit for lifetime of a compound nucleus.[33]
  6. ^ 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.[35] 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.[36]
  7. ^ Not all decay modes are caused by electrostatic repulsion. For example, beta decay is caused by the weak interaction.[43]
  8. ^ 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.[48]
  9. ^ 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.[53] The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[54] 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).[55]
  10. ^ 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).[44] 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.
  11. Georgy Flerov,[56] a leading scientist at JINR, and thus it was a "hobbyhorse" for the facility.[57] 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.[33] They thus preferred to link new isotopes to the already known ones by successive alpha decays.[56]
  12. ^ For instance, element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm, Stockholm County, Sweden.[58] 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.[59] 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.[59] JINR insisted that they were the first to create the element and suggested a name of their own for the new element, joliotium;[60] the Soviet name was also not accepted (JINR later referred to the naming of the element 102 as "hasty").[61] This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements, signed 29 September 1992.[61] The name "nobelium" remained unchanged on account of its widespread usage.[62]
  13. ^ In Russian, Oganessian's name is spelled Оганесян [ˈɐgənʲɪˈsʲan]; the transliteration in accordance with the rules of the English language would be Oganesyan, with one s. Similarly, the Russian name for the element is оганесон, letter-for-letter oganeson. Oganessian is the Russified version of the Armenian last name Hovhannisyan (Armenian: Հովհաննիսյան [hɔvhɑnnisˈjɑn]). It means "son of Hovhannes", i.e., "son of John". It is the
    most common surname in Armenia
    .

References

  1. The Periodic Table of Videos
    . University of Nottingham. 15 December 2016.
  2. ^ Ritter, Malcolm (9 June 2016). "Periodic table elements named for Moscow, Japan, Tennessee". Associated Press. Retrieved 19 December 2017.
  3. ^
    PMID 16833687
    .
  4. ^ .
  5. ^ .
  6. ^ .
  7. ^ . Retrieved 18 January 2008.
  8. ^ .
  9. ^ Oganesson, American Elements
  10. ^ Oganesson - Element information, properties and uses, Royal Chemical Society
  11. ^ .
  12. .
  13. ^ . Retrieved 18 January 2008.
  14. .
  15. IUPAC. 30 November 2016. Archived
    from the original on 30 November 2016. Retrieved 1 December 2015.
  16. ^ St. Fleur, Nicholas (1 December 2016). "Four New Names Officially Added to the Periodic Table of Elements". The New York Times. Retrieved 1 December 2016.
  17. ^
    IUPAC. 8 June 2016. Archived
    from the original on 8 June 2016.
  18. ^ Hoffman, Ghiorso & Seaborg 2000, pp. 187–189.
  19. ^
    PMID 32959952
    .
  20. ^ Krämer, K. (2016). "Explainer: superheavy elements". Chemistry World. Retrieved 15 March 2020.
  21. ^ "Discovery of Elements 113 and 115". Lawrence Livermore National Laboratory. Archived from the original on 11 September 2015. Retrieved 15 March 2020.
  22. S2CID 127060181
    .
  23. .
  24. S2CID 123288075. Archived from the original
    (PDF) on 7 June 2015. Retrieved 20 October 2012.
  25. ^ Subramanian, S. (28 August 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 18 January 2020.
  26. ^ 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.
  27. ^ Hinde, D. (2017). "Something new and superheavy at the periodic table". The Conversation. Retrieved 30 January 2020.
  28. .
  29. .
  30. .
  31. ^ .
  32. .
  33. ^ .
  34. ^ a b c d Chemistry World (2016). "How to Make Superheavy Elements and Finish the Periodic Table [Video]". Scientific American. Retrieved 27 January 2020.
  35. ^ Hoffman, Ghiorso & Seaborg 2000, p. 334.
  36. ^ Hoffman, Ghiorso & Seaborg 2000, p. 335.
  37. ^ Zagrebaev, Karpov & Greiner 2013, p. 3.
  38. ^ Beiser 2003, p. 432.
  39. ^ 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.
  40. ^ 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.
  41. ISSN 0556-2813
    .
  42. ^ Audi et al. 2017, pp. 030001-129–030001-138.
  43. ^ Beiser 2003, p. 439.
  44. ^ a b Beiser 2003, p. 433.
  45. ^ Audi et al. 2017, p. 030001-125.
  46. S2CID 125849923
    .
  47. ^ Beiser 2003, p. 432–433.
  48. ^
    ISSN 1742-6596
    .
  49. ^ 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.
  50. ^ . Retrieved 16 February 2020.
  51. .
  52. .
  53. .
  54. .
  55. ^ Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table". Chemical & Engineering News. Retrieved 27 January 2020.
  56. ^
    Distillations
    . Retrieved 22 February 2020.
  57. ^ "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.
  58. ^ "Nobelium - Element information, properties and uses | Periodic Table". Royal Society of Chemistry. Retrieved 1 March 2020.
  59. ^ a b Kragh 2018, pp. 38–39.
  60. ^ Kragh 2018, p. 40.
  61. ^ (PDF) from the original on 25 November 2013. Retrieved 7 September 2016.
  62. .
  63. ^ Kragh 2018, p. 6.
  64. ^ Leach, Mark R. "The INTERNET Database of Periodic Tables". Retrieved 8 July 2016.
  65. ^ . Retrieved 4 October 2013.
  66. .
  67. ^ .
  68. )
  69. .
  70. ^ Public Affairs Department, Lawrence Berkeley Laboratory (21 July 2001). "Results of element 118 experiment retracted". Archived from the original on 29 January 2008. Retrieved 18 January 2008.
  71. S2CID 4398009
    .
  72. ^ "Element 118 disappears two years after it was discovered". Physics World. 2 August 2001. Retrieved 2 April 2012.
  73. ^ Zagrebaev, Karpov & Greiner 2013.
  74. ^ Oganessian, Yu. T.; et al. (2002). "Results from the first 249
    Cf
    +48
    Ca
    experiment"
    (PDF). JINR Communication. Archived from the original (PDF) on 13 December 2004. Retrieved 13 June 2009.
  75. ^ .
  76. ^ Oganessian, Yu. T.; et al. (2002). "Element 118: results from the first 249
    Cf
    + 48
    Ca
    experiment"
    . Communication of the Joint Institute for Nuclear Research. Archived from the original on 22 July 2011.
  77. ^ "Livermore scientists team with Russia to discover element 118". Livermore press release. 3 December 2006. Archived from the original on 17 October 2011. Retrieved 18 January 2008.
  78. S2CID 55782333
    .
  79. .
  80. ^ Schewe, P. & Stein, B. (17 October 2006). "Elements 116 and 118 Are Discovered". Physics News Update. American Institute of Physics. Archived from the original on 1 January 2012. Retrieved 18 January 2008.
  81. ^ Weiss, R. (17 October 2006). "Scientists Announce Creation of Atomic Element, the Heaviest Yet". The Washington Post. Retrieved 18 January 2008.
  82. .
  83. ^ "Oganesson". WebElements Periodic Table. Retrieved 19 August 2019.
  84. . Retrieved 18 January 2008. I would say we're very confident.
  85. ^ Discovery and Assignment of Elements with Atomic Numbers 113, 115, 117 and 118. IUPAC (30 December 2015)
  86. S2CID 102228960
    .
  87. ^ .
  88. ^ Sychev, Vladimir (8 February 2017). "Юрий Оганесян: мы хотим узнать, где кончается таблица Менделеева" [Yuri Oganessian: we want to know where the Mendeleev table ends]. RIA Novosti (in Russian). Retrieved 31 March 2017.
  89. ^ Roberto, J. B. (31 March 2015). "Actinide Targets for Super-Heavy Element Research" (PDF). cyclotron.tamu.edu. Texas A & M University. Retrieved 28 April 2017.
  90. ^ Hauschild, K. (26 June 2019). Superheavy nuclei at RIKEN, Dubna, and JYFL (PDF). Conseil Scientifique de l'IN2P3. Retrieved 31 July 2019.
  91. ^ Hauschild, K. (2019). Heavy nuclei at RIKEN, Dubna, and JYFL (PDF). Conseil Scientifique de l'IN2P3. Retrieved 1 August 2019.
  92. .
  93. .
  94. ^ "Discovery of New Elements Makes Front Page News". Berkeley Lab Research Review Summer 1999. 1999. Retrieved 18 January 2008.
  95. ^ Koppenol, W. H. (2002). "Naming of new elements (IUPAC Recommendations 2002)" (PDF).
    S2CID 95859397
    .
  96. ^ "New chemical elements discovered in Russia's Science City". 12 February 2007. Retrieved 9 February 2008.
  97. ^ Yemel'yanova, Asya (17 December 2006). "118-й элемент назовут по-русски (118th element will be named in Russian)" (in Russian). vesti.ru. Archived from the original on 25 December 2008. Retrieved 18 January 2008.
  98. ^ "Российские физики предложат назвать 116 химический элемент московием (Russian Physicians Will Suggest to Name Element 116 Moscovium)" (in Russian). rian.ru. 2011. Retrieved 8 May 2011.
  99. ^ "News: Start of the Name Approval Process for the Elements of Atomic Number 114 and 116". International Union of Pure and Applied Chemistry. Archived from the original on 23 August 2014. Retrieved 2 December 2011.
  100. S2CID 95859397
    .
  101. .
  102. ^ "What it takes to make a new element". Chemistry World. Retrieved 3 December 2016.
  103. ^ a b Tarasevich, Grigoriy; Lapenko, Igor (2019). "Юрий Оганесян о тайнах ядра, новых элементах и смысле жизни" [Yuri Oganessian about the secret of the nucleus, new elements and the meaning of life]. Kot Shryodingyera (in Russian). No. Special. Direktsiya Festivalya Nauki. p. 22.
  104. ^ a b Gray, Richard (11 April 2017). "Mr Element 118: The only living person on the periodic table". New Scientist. Retrieved 26 April 2017.
  105. ^ 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.
  106. ^ Subramanian, S. "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 18 January 2020.
  107. S2CID 4415582
    .
  108. .
  109. .
  110. .
  111. ^ "Oganesson - Element information, properties and uses | Periodic Table". www.rsc.org. Retrieved 25 January 2023.
  112. ^ "Oganesson - Protons - Neutrons - Electrons - Electron Configuration". Material Properties. 8 December 2020. Retrieved 25 January 2023.
  113. ^
    S2CID 118739116
    .
  114. .
  115. .
  116. .
  117. ^ .
  118. .
  119. .
  120. .
  121. ^ Bader, Richard 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.
  122. ^ "Ununoctium (Uuo) – Chemical properties, Health and Environmental effects". Lenntech. Archived from the original on 16 January 2008. Retrieved 18 January 2008.
  123. ^ .
  124. .
  125. . Retrieved 15 September 2015.
  126. (PDF) on 15 January 2018. Retrieved 15 January 2018.
  127. .
  128. .
  129. ^ .
  130. .
  131. ^ Garisto, Dan (12 February 2018). "5 ways the heaviest element on the periodic table is really bizarre". ScienceNews. Retrieved 12 February 2023.
  132. PMID 31343819
    .
  133. ^ "Oganesson: Compounds Information". WebElements Periodic Table. Retrieved 19 August 2019.
  134. ^ .
  135. .
  136. .
  137. .
  138. ^ a b Seaborg, Glenn Theodore (c. 2006). "transuranium element (chemical element)". Britannica Online. Retrieved 16 March 2010.
  139. S2CID 235259897
    . Retrieved 30 June 2021.

Bibliography

Further reading

External links