Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Nihonium

Nihonium is a synthetic chemical element; it has symbol Nh


Nihonium, 113Nh
and atomic number 113. It is extremely radioactive: its most
stable known isotope, nihonium-286, has a half-life of about Nihonium
10 seconds. In the periodic table, nihonium is a transactinide Pronunciation /nɪˈhoʊniəm/ ​
element in the p-block. It is a member of period 7 and group (nih-HOH-nee-əm)
13. Mass number [286]

Nihonium was first reported to have been created in Nihonium in the periodic table
experiments being carried out between 14 July to August 10, Tl
2003, by a Russian–American collaboration at the Joint ↑
Nh
Institute for Nuclear Research (JINR) in Dubna, Russia, ↓

working in collaboration with the Lawrence Livermore copernicium ← nihonium → flerovium
National Laboratory in Livermore, California,[11] and in
July 23, 2004, by a team of Japanese scientists at Riken in Atomic number (Z) 113
Wakō, Japan.[12] The confirmation of their claims in the Group group 13 (boron
ensuing years involved independent teams of scientists group)
working in the United States, Germany, Sweden, and China, Period period 7
as well as the original claimants in Russia and Japan. In
Block p-block
2015, the IUPAC/IUPAP Joint Working Party recognized
the element and assigned the priority of the discovery and Electron [Rn] 5f14 6d10 7s2
configuration 7p1 (predicted)[1]
naming rights for the element to Riken.[13] The Riken team
suggested the name nihonium in 2016, which was approved Electrons per shell 2, 8, 18, 32, 32, 18,
in the same year. The name comes from the common 3 (predicted)
Japanese name for Japan ( 日本 , nihon). Physical properties
Phase at STP solid
Very little is known about nihonium, as it has only been
(predicted)[1][2][3]
made in very small amounts that decay within seconds. The
anomalously long lives of some superheavy nuclides, Melting point 700 K ​(430 °C, ​
including some nihonium isotopes, are explained by the 810 °F)
"island of stability" theory. Experiments support the theory, (predicted)[1]
with the half-lives of the confirmed nihonium isotopes Boiling point 1430 K ​(1130 °C, ​
increasing from milliseconds to seconds as neutrons are 2070 °F)
added and the island is approached. Nihonium has been (predicted)[1][4]
calculated to have similar properties to its homologues Density (near r.t.) 16 g/cm3
boron, aluminium, gallium, indium, and thallium. All but (predicted)[4]
boron are post-transition metals, and nihonium is expected to Heat of fusion 7.61 kJ/mol
be a post-transition metal as well. It should also show several
(extrapolated)[3]
major differences from them; for example, nihonium should
Heat of 130 kJ/mol
be more stable in the +1 oxidation state than the +3 state,
vaporisation (predicted)[2][4]
like thallium, but in the +1 state nihonium should behave
Atomic properties
more like silver and astatine than thallium. Preliminary Oxidation states (−1), (+1), (+3),
experiments in 2017 showed that elemental nihonium is not (+5)
very volatile; its chemistry remains largely unexplored. (predicted)[1][4][5]
Ionisation 1st: 704.9 kJ/mol
Introduction energies (predicted)[1]
2nd: 2240 kJ/mol
(predicted)[4]
Synthesis of superheavy nuclei
3rd: 3020 kJ/mol
(predicted)[4]
(more)

Atomic radius empirical: 170 pm


(predicted)[1]
Covalent radius 172–180 pm
(extrapolated)[3]
Other properties
Natural synthetic
occurrence
Crystal structure ​ exagonal close-
h
packed (hcp)
A graphic depiction of a nuclear
fusion reaction. Two nuclei fuse into
one, emitting a neutron. Reactions
(predicted)[6][7]
that created new elements to this
moment were similar, with the only CAS Number 54084-70-7
possible difference that several History
singular neutrons sometimes were
released, or none at all. Naming After Japan (Nihon
in Japanese)
Discovery Riken (Japan, first
A superheavy[a] atomic nucleus is created in a nuclear
undisputed claim
reaction that combines two other nuclei of unequal size[b]
2004)
into one; roughly, the more unequal the two nuclei in terms
JINR (Russia) and
of mass, the greater the possibility that the two react.[19] The
Livermore (US, first
material made of the heavier nuclei is made into a target,
announcement
which is then bombarded by the beam of lighter nuclei. Two
2003)
nuclei can only fuse into one if they approach each other
closely enough; normally, nuclei (all positively charged) Isotopes of nihonium
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.[20] 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.[20]
Coming close enough alone is not enough for two nuclei to
Main isotopes[8] Decay
fuse: when two nuclei approach each other, they usually
remain together for about 10−20 second and then part ways abun­‐ half-life mode pro­‐

(not necessarily in the same composition as before the dance (t1/2) duct

reaction) rather than form a single nucleus.[20][21] This 278


Nh synth 0.002 s α 274
Rg
happens because during the attempted formation of a single 282 278
Nh synth 0.061 s α Rg
nucleus, electrostatic repulsion tears apart the nucleus that is
being formed.[20] Each pair of a target and a beam is 283
Nh synth 0.123 s α 279
Rg
characterized by its cross section—the probability that fusion 284 280
will occur if two nuclei approach one another expressed in Nh synth 0.90 s α Rg

terms of the transverse area that the incident particle must hit ε 284
Cn
in order for the fusion to occur.[c] This fusion may occur as a
285 281
result of the quantum effect in which nuclei can tunnel Nh synth 2.1 s α Rg

through electrostatic repulsion. If the two nuclei can stay SF –


close for past that phase, multiple nuclear interactions result
286 282
in redistribution of energy and an energy equilibrium.[20] Nh synth 9.5 s α Rg

287
Nh synth 5.5 s?[9] α 283
Rg
External videos
290 286
Visualization (https://www.youtub Nh synth 2 s?[10] α Rg

e.com/watch?v=YovAFlzFtzg) of
The resulting merger is an excited state[24]—termed a compound
unsuccessful nuclear fusion, based
on calculations from the Australian
nucleus—and thus it is very unstable.[20] To reach a more stable
National University[23]
state, the temporary merger may fission without formation of a
more stable nucleus.[25] 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 about 10−16 second after the initial nuclear collision and results in creation of a more stable
nucleus.[25] 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.[26][d]

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.[28] In the separator, the newly produced nucleus is separated from
other nuclides (that of the original beam and any other reaction products)[e] and transferred to a surface-
barrier detector, which stops the nucleus. The exact location of the upcoming impact on the detector is
marked; also marked are its energy and the time of the arrival.[28] The transfer takes about 10−6 second; in
order to be detected, the nucleus must survive this long.[31] The nucleus is recorded again once its decay is
registered, and the location, the energy, and the time of the decay are measured.[28]

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.[32] 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.[33][34] Superheavy nuclei are thus
theoretically predicted[35] and have so far been observed[36] to predominantly decay via decay modes that
are caused by such repulsion: alpha decay and spontaneous fission.[f] Almost all alpha emitters have over
210 nucleons,[38] and the lightest nuclide primarily undergoing spontaneous fission has 238.[39] In both
decay modes, nuclei are inhibited from decaying by corresponding energy barriers for each mode, but they
can be tunneled through.[33][34]

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.[41] Spontaneous fission
is caused by electrostatic Scheme of an apparatus for creation of superheavy elements, based on
repulsion tearing the nucleus the Dubna Gas-Filled Recoil Separator set up in the Flerov Laboratory of
apart and produces various Nuclear Reactions in JINR. The trajectory within the detector and the beam
nuclei in different instances of focusing apparatus changes because of a dipole magnet in the former and
identical nuclei fissioning.[34] As quadrupole magnets in the latter.[40]

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),[42] and by 30 orders of
magnitude from thorium (element 90) to fermium (element 100).[43] 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.[34][44] 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.[34][44] 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.[45] Experiments on lighter superheavy nuclei,[46] as well as those closer to the expected
island,[42] have shown greater than previously anticipated stability against spontaneous fission, showing the
importance of shell effects on nuclei.[g]

Alpha decays are registered by the emitted alpha particles, and the decay products are easy to determine
before the actual decay; if such a decay or a series of consecutive decays produces a known nucleus, the
original product of a reaction can be easily determined.[h] (That all decays within a decay chain were indeed
related to each other is established by the location of these decays, which must be in the same place.)[28]
The known nucleus can be recognized by the specific characteristics of decay it undergoes such as decay
energy (or more specifically, the kinetic energy of the emitted particle).[i] Spontaneous fission, however,
produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[j]

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

Early indications
The syntheses of elements 107 to 112 were conducted at the GSI Helmholtz Centre for Heavy Ion Research
in Darmstadt, Germany, from 1981 to 1996. These elements were made by cold fusion[l] reactions, in
which targets made of thallium, lead, and bismuth, which are around the stable configuration of 82 protons,
are bombarded with heavy ions of period 4 elements. This creates fused nuclei with low excitation energies
due to the stability of the targets' nuclei, significantly increasing the yield of superheavy elements. Cold
fusion was pioneered by Yuri Oganessian and his team in 1974 at the Joint Institute for Nuclear Research
(JINR) in Dubna, Soviet Union. Yields from cold fusion reactions were found to decrease significantly with
increasing atomic number; the resulting nuclei were severely neutron-deficient and short-lived. The GSI
team attempted to synthesise element 113 via cold fusion in 1998 and 2003, bombarding bismuth-209 with
zinc-70; both attempts were unsuccessful.[60][61]

Faced with this problem, Oganessian and his team at the JINR turned their renewed attention to the older
hot fusion technique, in which heavy actinide targets were bombarded with lighter ions. Calcium-48 was
suggested as an ideal projectile, because it is very neutron-rich for a light element (combined with the
already neutron-rich actinides) and would minimise the neutron deficiencies of the nuclides produced.
Being doubly magic, it would confer benefits in stability to the fused nuclei. In collaboration with the team
at the Lawrence Livermore National Laboratory (LLNL) in Livermore, California, United States, they
made an attempt on element 114 (which was predicted to be a magic number, closing a proton shell, and
more stable than element 113).[60]

In 1998, the JINR–LLNL collaboration started their attempt on element 114, bombarding a target of
plutonium-244 with ions of calcium-48:[60]

244 48
94Pu + 20Ca → 292114* → 290114 + 2 n + e− → 290113 + νe ?

A single atom was observed which was thought to be the isotope 289 114: the results were published in
January 1999.[62] Despite numerous attempts to repeat this reaction, an isotope with these decay properties
has never again been found, and the exact identity of this activity is unknown.[63] A 2016 paper by Sigurd
Hofmann et al. considered that the most likely explanation of the 1998 result is that two neutrons were
emitted by the produced compound nucleus, leading to 290 114 and electron capture to 290 113, while more
neutrons were emitted in all other produced chains. This would have been the first report of a decay chain
from an isotope of element 113, but it was not recognised at the time, and the assignment is still
uncertain.[10] A similar long-lived activity observed by the JINR team in March 1999 in the 242 Pu + 48 Ca
reaction may be due to the electron-capture daughter of 287 114, 287 113; this assignment is also tentative.[9]

JINR–LLNL collaboration
The now-confirmed discovery of element 114 was made in June 1999 when the JINR team repeated the
first 244 Pu + 48 Ca reaction from 1998;[64][65] following this, the JINR team used the same hot fusion
technique to synthesize elements 116 and 118 in 2000 and 2002 respectively via the 248 Cm + 48 Ca and
249 Cf + 48 Ca reactions. They then turned their attention to the missing odd-numbered elements, as the odd

protons and possibly neutrons would hinder decay by spontaneous fission and result in longer decay
chains.[60][66]
The first report of element 113 was in August 2003, when it was identified as an alpha decay product of
element 115. Element 115 had been produced by bombarding a target of americium-243 with calcium-48
projectiles. The JINR–LLNL collaboration published its results in February 2004:[66]

243 48
95Am + 20Ca → 291115* → 288115 + 3 n → 284113 + α
243 48
95Am + 20Ca → 291115* → 287115 + 4 n → 283113 + α

Four further alpha decays were observed, ending with the spontaneous fission of isotopes of element 105,
dubnium.[66]

Riken
While the JINR–LLNL collaboration had been studying fusion reactions with 48 Ca, a team of Japanese
scientists at the Riken Nishina Center for Accelerator-Based Science in Wakō, Japan, led by Kōsuke Morita
had been studying cold fusion reactions. Morita had previously studied the synthesis of superheavy
elements at the JINR before starting his own team at Riken. In 2001, his team confirmed the GSI's
discoveries of elements 108, 110, 111, and 112. They then made a new attempt on element 113, using the
same 209 Bi + 70 Zn reaction that the GSI had attempted unsuccessfully in 1998. Despite the much lower
yield expected than for the JINR's hot fusion technique with calcium-48, the Riken team chose to use cold
fusion as the synthesised isotopes would alpha decay to known daughter nuclides and make the discovery
much more certain, and would not require the use of radioactive targets.[67] In particular, the isotope 278 113
expected to be produced in this reaction would decay to the known 266 Bh, which had been synthesised in
2000 by a team at the Lawrence Berkeley National Laboratory (LBNL) in Berkeley.[68]

The bombardment of 209 Bi with 70 Zn at Riken began in September 2003.[69] The team detected a single
atom of 278 113 in July 2004 and published their results that September:[70]

209
83Bi + 70
30Zn →
279113* → 278113 + n

The Riken team observed four alpha decays from 278 113, creating a decay chain passing through 274 Rg,
270 Mt, and 266 Bh before terminating with the spontaneous fission of 262 Db.[70] The decay data they

observed for the alpha decay of 266 Bh matched the 2000 data, lending support for their claim. Spontaneous
fission of its daughter 262 Db had not been previously known; the American team had observed only alpha
decay from this nuclide.[68]

Road to confirmation
When the discovery of a new element is claimed, the Joint Working Party (JWP) of the International Union
of Pure and Applied Chemistry (IUPAC) and the International Union of Pure and Applied Physics (IUPAP)
assembles to examine the claims according to their criteria for the discovery of a new element, and decides
scientific priority and naming rights for the elements. According to the JWP criteria, a discovery must
demonstrate that the element has an atomic number different from all previously observed values. It should
also preferably be repeated by other laboratories, although this requirement has been waived where the data
is of very high quality. Such a demonstration must establish properties, either physical or chemical, of the
new element and establish that they are those of a previously unknown element. The main techniques used
to demonstrate atomic number are cross-reactions (creating claimed nuclides as parents or daughters of other
nuclides produced by a different reaction) and anchoring decay chains to known daughter nuclides. For the
JWP, priority in confirmation takes precedence over the date of the original claim. Both teams set out to
confirm their results by these methods.[71]

Summary of decay chains passing through isotopes of element 113, ending at mendelevium (element 101) or
earlier. The two chains with bold-bordered nuclides were accepted by the JWP as evidence for the discoveries of
element 113 and its parents, elements 115 and 117. Data is presented as known in 2015 (before the JWP's
conclusions were published).

2004–2008
In June 2004 and again in December 2005, the JINR–LLNL collaboration strengthened their claim for the
discovery of element 113 by conducting chemical experiments on 268 Db, the final decay product of 288 115.
This was valuable as none of the nuclides in this decay chain were previously known, so that their claim
was not supported by any previous experimental data, and chemical experimentation would strengthen the
case for their claim, since the chemistry of dubnium is known. 268 Db was successfully identified by
extracting the final decay products, measuring spontaneous fission (SF) activities and using chemical
identification techniques to confirm that they behave like a group 5 element (dubnium is known to be in
group 5).[1][72] Both the half-life and decay mode were confirmed for the proposed 268 Db which lends
support to the assignment of the parent and daughter nuclei to elements 115 and 113 respectively.[72][73]
Further experiments at the JINR in 2005 confirmed the observed decay data.[68]
In November and December 2004, the Riken team studied the 205 Tl + 70 Zn reaction, aiming the zinc beam
onto a thallium rather than a bismuth target, in an effort to directly produce 274 Rg in a cross-bombardment
as it is the immediate daughter of 278 113. The reaction was unsuccessful, as the thallium target was
physically weak compared to the more commonly used lead and bismuth targets, and it deteriorated
significantly and became non-uniform in thickness. The reasons for this weakness are unknown, given that
thallium has a higher melting point than bismuth.[74] The Riken team then repeated the original 209 Bi +
70 Zn reaction and produced a second atom of 278 113 in April 2005, with a decay chain that again

terminated with the spontaneous fission of 262 Db. The decay data were slightly different from those of the
first chain: this could have been because an alpha particle escaped from the detector without depositing its
full energy, or because some of the intermediate decay products were formed in metastable isomeric
states.[68]

In 2006, a team at the Heavy Ion Research Facility in Lanzhou, China, investigated the 243 Am + 26 Mg
reaction, producing four atoms of 266 Bh. All four chains started with an alpha decay to 262 Db; three chains
ended there with spontaneous fission, as in the 278 113 chains observed at Riken, while the remaining one
continued via another alpha decay to 258 Lr, as in the 266 Bh chains observed at LBNL.[71]

In June 2006, the JINR–LLNL collaboration claimed to have synthesised a new isotope of element 113
directly by bombarding a neptunium-237 target with accelerated calcium-48 nuclei:

237
93Np + 48
20Ca →
285113* → 282113 + 3 n

Two atoms of 282 113 were detected. The aim of this experiment had been to synthesise the isotopes 281 113
and 282 113 that would fill in the gap between isotopes produced via hot fusion (283 113 and 284 113) and
cold fusion (278 113). After five alpha decays, these nuclides would reach known isotopes of lawrencium,
assuming that the decay chains were not terminated prematurely by spontaneous fission. The first decay
chain ended in fission after four alpha decays, presumably originating from 266 Db or its electron-capture
daughter 266 Rf. Spontaneous fission was not observed in the second chain even after four alpha decays. A
fifth alpha decay in each chain could have been missed, since 266 Db can theoretically undergo alpha decay,
in which case the first decay chain would have ended at the known 262 Lr or 262 No and the second might
have continued to the known long-lived 258 Md, which has a half-life of 51.5 days, longer than the duration
of the experiment: this would explain the lack of a spontaneous fission event in this chain. In the absence of
direct detection of the long-lived alpha decays, these interpretations remain unconfirmed, and there is still no
known link between any superheavy nuclides produced by hot fusion and the well-known main body of the
chart of nuclides.[75]

2009–2015
The JWP published its report on elements 113–116 and 118 in 2011. It recognised the JINR–LLNL
collaboration as having discovered elements 114 and 116, but did not accept either team's claim to element
113 and did not accept the JINR–LLNL claims to elements 115 and 118. The JINR–LLNL claim to
elements 115 and 113 had been founded on chemical identification of their daughter dubnium, but the JWP
objected that current theory could not distinguish between superheavy group 4 and group 5 elements by
their chemical properties with enough confidence to allow this assignment.[68] The decay properties of all
the nuclei in the decay chain of element 115 had not been previously characterised before the JINR
experiments, a situation which the JWP generally considers "troublesome, but not necessarily exclusive",
and with the small number of atoms produced with neither known daughters nor cross-reactions the JWP
considered that their criteria had not been fulfilled.[68] The JWP did not accept the Riken team's claim either
due to inconsistencies in the decay data, the small number of atoms of element 113 produced, and the lack
of unambiguous anchors to known isotopes.[68]

In early 2009, the Riken team synthesised the decay product 266 Bh directly in the 248 Cm + 23 Na reaction
to establish its link with 278 113 as a cross-bombardment. They also established the branched decay of
262 Db, which sometimes underwent spontaneous fission and sometimes underwent the previously known

alpha decay to 258 Lr.[76][77]

In late 2009, the JINR–LLNL collaboration studied the 249 Bk + 48 Ca reaction in an effort to produce
element 117, which would decay to elements 115 and 113 and bolster their claims in a cross-reaction. They
were now joined by scientists from Oak Ridge National Laboratory (ORNL) and Vanderbilt University,
both in Tennessee, United States,[60] who helped procure the rare and highly radioactive berkelium target
necessary to complete the JINR's calcium-48 campaign to synthesise the heaviest elements on the periodic
table.[60] Two isotopes of element 117 were synthesised, decaying to element 115 and then element 113:[78]

249
97Bk + 48
20Ca →
297117* → 294117 + 3 n → 290115 + α → 286113 + α
249 48
97Bk + 20Ca → 297117* → 293117 + 4 n → 289115 + α → 285113 + α

The new isotopes 285 113 and 286 113 produced did not overlap with the previously claimed 282 113, 283 113,
and 284 113, so this reaction could not be used as a cross-bombardment to confirm the 2003 or 2006
claims.[71]

In March 2010, the Riken team again attempted to synthesise 274 Rg directly through the 205 Tl + 70 Zn
reaction with upgraded equipment; they failed again and abandoned this cross-bombardment route.[74]

After 450 more days of irradiation of bismuth with zinc projectiles, Riken produced and identified another
278 113 atom in August 2012.[79] Although electricity prices had soared since the
2011 Tōhoku earthquake
and tsunami, and Riken had ordered the shutdown of the accelerator programs to save money, Morita's team
was permitted to continue with one experiment, and they chose their attempt to confirm their synthesis of
element 113.[80] In this case, a series of six alpha decays was observed, leading to an isotope of
mendelevium:

278113 274 270 266 262 258 254


→ 111Rg + α → 109Mt + α → 107Bh + α → 105Db + α → 103Lr + α → 101Md + α

This decay chain differed from the previous observations at Riken mainly in the decay mode of 262 Db,
which was previously observed to undergo spontaneous fission, but in this case instead alpha decayed; the
alpha decay of 262 Db to 258 Lr is well-known. The team calculated the probability of accidental coincidence
to be 10−28 , or totally negligible.[79] The resulting 254 Md atom then underwent electron capture to 254 Fm,
which underwent the seventh alpha decay in the chain to the long-lived 250 Cf, which has a half-life of
around thirteen years.[81]

The 249 Bk + 48 Ca experiment was repeated at the JINR in 2012 and 2013 with consistent results, and
again at the GSI in 2014.[71] In August 2013, a team of researchers at Lund University in Lund, Sweden,
and at the GSI announced that they had repeated the 2003 243 Am + 48 Ca experiment, confirming the
findings of the JINR–LLNL collaboration.[69][82] The same year, the 2003 experiment had been repeated at
the JINR, now also creating the isotope 289 115 that could serve as a cross-bombardment for confirming
their discovery of the element 117 isotope 293 117, as well as its daughter 285 113 as part of its decay
chain.[71] Confirmation of 288 115 and its daughters was published by the team at the LBNL in August
2015.[83]

Approval of discoveries
In December 2015, the conclusions of a new JWP report were published by IUPAC in a press release, in
which element 113 was awarded to Riken; elements 115, 117, and 118 were awarded to the collaborations
involving the JINR.[84] A joint 2016 announcement by IUPAC and IUPAP had been scheduled to coincide
with the publication of the JWP reports, but IUPAC alone decided on an early release because the news of
Riken being awarded credit for element 113 had been leaked to Japanese newspapers.[85] For the first time
in history, a team of Asian physicists would name a new element.[84] The JINR considered the awarding of
element 113 to Riken unexpected, citing their own 2003 production of elements 115 and 113, and pointing
to the precedents of elements 103, 104, and 105 where IUPAC had awarded joint credit to the JINR and
LBNL. They stated that they respected IUPAC's decision, but reserved determination of their position for
the official publication of the JWP reports.[86]

The full JWP reports were published on 21 January 2016. The JWP recognised the discovery of element
113, assigning priority to Riken. They noted that while the individual decay energies of each nuclide in the
decay chain of 278 113 were inconsistent, their sum was now confirmed to be consistent, strongly suggesting
that the initial and final states in 278 113 and its daughter 262 Db were the same for all three events. The
decay of 262 Db to 258 Lr and 254 Md was previously known, firmly anchoring the decay chain of 278 113 to
known regions of the chart of nuclides. The JWP considered that the JINR–LLNL collaborations of 2004
and 2007, producing element 113 as the daughter of element 115, did not meet the discovery criteria as they
had not convincingly determined the atomic numbers of their nuclides through cross-bombardments, which
were considered necessary since their decay chains were not anchored to previously known nuclides. They
also considered that the previous JWP's concerns over their chemical identification of the dubnium daughter
had not been adequately addressed. The JWP recognised the JINR–LLNL–ORNL–Vanderbilt
collaboration of 2010 as having discovered elements 117 and 115, and accepted that element 113 had been
produced as their daughter, but did not give this work shared credit.[71][74][87]

After the publication of the JWP reports, Sergey Dimitriev, the lab director of the Flerov lab at the JINR
where the discoveries were made, remarked that he was happy with IUPAC's decision, mentioning the time
Riken spent on their experiment and their good relations with Morita, who had learnt the basics of
synthesising superheavy elements at the JINR.[60][86]

The sum argument advanced by the JWP in the approval of the discovery of element 113 was later criticised
in a May 2016 study from Lund University and the GSI, as it is only valid if no gamma decay or internal
conversion takes place along the decay chain, which is not likely for odd nuclei, and the uncertainty of the
alpha decay energies measured in the 278 113 decay chain was not small enough to rule out this possibility.
If this is the case, similarity in lifetimes of intermediate daughters becomes a meaningless argument, as
different isomers of the same nuclide can have different half-lives: for example, the ground state of 180 Ta
has a half-life of hours, but an excited state 180mTa has never been observed to decay. This study found
reason to doubt and criticise the IUPAC approval of the discoveries of elements 115 and 117, but the data
from Riken for element 113 was found to be congruent, and the data from the JINR team for elements 115
and 113 to probably be so, thus endorsing the IUPAC approval of the discovery of element 113.[88][89]
Two members of the JINR team published a journal article rebutting these criticisms against the congruence
of their data on elements 113, 115, and 117 in June 2017.[90]

Naming
Using Mendeleev's nomenclature for unnamed and undiscovered
elements, nihonium would be known as eka-thallium. In 1979,
IUPAC published recommendations according to which the
element was to be called ununtrium (with the corresponding symbol
of Uut),[91] a systematic element name as a placeholder, until the
discovery of the element is confirmed and a name is decided on.
The recommendations were widely used in the chemical
community on all levels, from chemistry classrooms to advanced
textbooks, but were mostly ignored among scientists in the field, Kōsuke Morita and Hiroshi
who called it "element 113", with the symbol of E113, (113), or Matsumoto, celebrating the naming
even simply 113.[1] on 1 December 2016.

Before the JWP recognition of their priority, the Japanese team had
unofficially suggested various names: japonium, after their home country;[92] nishinanium, after Japanese
physicist Yoshio Nishina, the "founding father of modern physics research in Japan";[93] and rikenium, after
the institute.[92] After the recognition, the Riken team gathered in February 2016 to decide on a name.
Morita expressed his desire for the name to honour the fact that element 113 had been discovered in Japan.
Japonium was considered, making the connection to Japan easy to identify for non-Japanese, but it was
rejected as Jap is considered an ethnic slur. The name nihonium was chosen after an hour of deliberation: it
comes from nihon ( 日本 ), one of the two Japanese pronunciations for the name of Japan.[94] The
discoverers also intended to reference the support of their research by the Japanese people (Riken being
almost entirely government-funded),[95] recover lost pride and trust in science among those who were
affected by the Fukushima Daiichi nuclear disaster,[96] and honour Japanese chemist Masataka Ogawa's
1908 discovery of rhenium, which he named "nipponium" with symbol Np after the other Japanese
pronunciation of Japan's name.[87] As Ogawa's claim had not been accepted, the name "nipponium" could
not be reused for a new element, and its symbol Np had since been used for neptunium.[m] In March 2016,
Morita proposed the name "nihonium" to IUPAC, with the symbol Nh.[87] The naming realised what had
been a national dream in Japanese science ever since Ogawa's claim.[80]

The former president of IUPAP, Cecilia Jarlskog, complained at the Nobel Symposium on Superheavy
Elements in Bäckaskog Castle, Sweden, in June 2016 about the lack of openness involved in the process of
approving new elements, and stated that she believed that the JWP's work was flawed and should be redone
by a new JWP. A survey of physicists determined that many felt that the Lund–GSI 2016 criticisms of the
JWP report were well-founded, but that the conclusions would hold up if the work was redone, and the new
president, Bruce McKellar, ruled that the proposed names should be released in a joint IUPAP–IUPAC
press release.[85] Thus, IUPAC and IUPAP publicised the proposal of nihonium that June,[96] and set a
five-month term to collect comments, after which the name would be formally established at a
conference.[99][100] The name was officially approved in November 2016.[101] The naming ceremony for
the new element was held in Tokyo, Japan, in March 2017, with Naruhito, then the Crown Prince of Japan,
in attendance.[102]

Isotopes
Nihonium has no stable or naturally
occurring isotopes. Several radioactive List of nihonium isotopes
isotopes have been synthesised in the
Half-life[n] Decay Discovery Discovery
laboratory, either by fusing two atoms Isotope
mode year reaction
or by observing the decay of heavier Value ref
elements. Eight different isotopes of 278
Nh 2.3 ms [103] α 2004 209
Bi(70Zn,n)
nihonium have been reported with 282 [104] 237
Nh 61 ms α 2006 Np(48Ca,3n)
atomic masses 278, 282–287, and 290
(287 Nh and 290 Nh are unconfirmed); 283
Nh 123 ms [104] α 2004 287
Mc(—,α)
they all decay through alpha decay to 284
Nh 0.90 s [104] α, EC 2004 288
Mc(—,α)
isotopes of roentgenium.[106] There
285 2.1 s [104] α, SF 2010 289
have been indications that nihonium- Nh Mc(—,α)

284 can also decay by electron capture 286


Nh 9.5 s [105] α 2010 290
Mc(—,α)
to copernicium-284, though estimates 287
287
Nh[o] 5.5 s [9] α 1999 Fl(e−,νe)
of the partial half-life for this branch
vary strongly by model.[107] A 290
Nh[o] 2s [10] α 1998 290
Fl(e−,νe)
spontaneous fission branch of
nihonium-285 has also been
reported.[104]

Stability and half-lives


The stability of nuclei quickly
decreases with the increase in atomic
number after curium, element 96,
whose half-life is over ten thousand
times 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: this is because of the
ever-increasing Coulomb repulsion of
A chart of heavy nuclides with their known and predicted half-lives
protons, so that the strong nuclear force (known nuclides shown with borders). Nihonium (row 113) is
cannot hold the nucleus together expected to be within the "island of stability" (white circle) and thus
against spontaneous fission for long. its nuclei are slightly more stable than would otherwise be predicted;
Calculations suggest that in the the known nihonium isotopes are too neutron-poor to be within the
absence of other stabilising factors, island.
elements with more than 103 protons
should not exist. Researchers in the
1960s suggested that the closed nuclear shells around 114 protons and 184 neutrons should counteract this
instability, and create an "island of stability" containing nuclides with half-lives reaching thousands or
millions of years. The existence of the island is still unproven, but the existence of the superheavy elements
(including nihonium) confirms that the stabilising effect is real, and in general the known superheavy
nuclides become longer-lived as they approach the predicted location of the island.[108][109]

All nihonium isotopes are unstable and radioactive; the heavier nihonium isotopes are more stable than the
lighter ones, as they are closer to the centre of the island. The most stable known nihonium isotope, 286 Nh,
is also the heaviest; it has a half-life of 8 seconds. The isotope 285 Nh, as well as the unconfirmed 287 Nh
and 290 Nh, have also been reported to have half-lives of over a second. The isotopes 284 Nh and 283 Nh
have half-lives of 0.90 and 0.12 seconds respectively. The remaining two isotopes have half-lives between
0.1 and 100 milliseconds: 282 Nh has a half-life of 61 milliseconds, and 278 Nh, the lightest known nihonium
isotope, is also the shortest-lived, with a half-life of 1.4 milliseconds. This rapid increase in the half-lives
near the closed neutron shell at N = 184 is seen in roentgenium, copernicium, and nihonium (elements 111
through 113), where each extra neutron so far multiplies the half-life by a factor of 5 to 20.[109][110]

Predicted properties
Very few properties of nihonium or its compounds have been measured; this is due to its extremely limited
and expensive production[111] and the fact it decays very quickly. Properties of nihonium mostly remain
unknown and only predictions are available.

Physical and atomic


Nihonium is the first member of the 7p series of elements and the
heaviest group 13 element on the periodic table, below boron,
aluminium, gallium, indium, and thallium. All the group 13
elements except boron are metals, and nihonium is expected to
follow suit. Nihonium is predicted to show many differences from
its lighter homologues. The major reason for this is the spin–orbit
(SO) interaction, which is especially strong for the superheavy
elements, because their electrons move much faster than in lighter
atoms, at velocities close to the speed of light.[5] In relation to
nihonium atoms, it lowers the 7s and the 7p electron energy levels
(stabilising those electrons), but two of the 7p electron energy levels
are stabilised more than the other four.[113] The stabilisation of the
7s electrons is called the inert pair effect, and the separation of the
7p subshell into the more and less stabilised parts is called subshell
splitting. Computational chemists see the split as a change of the
second, azimuthal quantum number l, from 1 to 1/2 and 3/2 for the Atomic energy levels of outermost s,
more and less stabilised parts of the 7p subshell, respectively.[5][p] p, and d electrons of thallium and
For theoretical purposes, the valence electron configuration may be nihonium[112]
represented to reflect the 7p subshell split as 7s2 7p1/2 1 .[1] The first
ionisation energy of nihonium is expected to be 7.306 eV, the
highest among the metals of group 13.[1] Similar subshell splitting should exist for the 6d electron levels,
with four being 6d3/2 and six being 6d5/2 . Both these levels are raised to be close in energy to the 7s ones,
high enough to possibly be chemically active. This would allow for the possibility of exotic nihonium
compounds without lighter group 13 analogues.[113]
Periodic trends would predict nihonium to have an atomic radius larger than that of thallium due to it being
one period further down the periodic table, but calculations suggest nihonium has an atomic radius of about
170 pm, the same as that of thallium, due to the relativistic stabilisation and contraction of its 7s and 7p1/2
orbitals. Thus, nihonium is expected to be much denser than thallium, with a predicted density of about 16
to 18 g/cm3 compared to thallium's 11.85 g/cm3 , since nihonium atoms are heavier than thallium atoms but
have the same volume.[1][112] Bulk nihonium is expected to have a hexagonal close-packed crystal
structure, like thallium.[6] The melting and boiling points of nihonium have been predicted to be 430 °C and
1100 °C respectively, exceeding the values for indium and thallium, following periodic trends.[1][2]
Nihonium should have a bulk modulus of 20.8 GPa, about half that of thallium (43 GPa).[7]

Chemical
The chemistry of nihonium is expected to be very different from that of thallium. This difference stems from
the spin–orbit splitting of the 7p shell, which results in nihonium being between two relatively inert closed-
shell elements (copernicium and flerovium).[114] Nihonium is expected to be less reactive than thallium,
because of the greater stabilisation and resultant chemical inactivity of the 7s subshell in nihonium
compared to the 6s subshell in thallium.[4] The standard electrode potential for the Nh+/Nh couple is
predicted to be 0.6 V. Nihonium should be a rather noble metal.[4]

The metallic group 13 elements are typically found in two oxidation states: +1 and +3. The former results
from the involvement of only the single p electron in bonding, and the latter results in the involvement of all
three valence electrons, two in the s-subshell and one in the p-subshell. Going down the group, bond
energies decrease and the +3 state becomes less stable, as the energy released in forming two additional
bonds and attaining the +3 state is not always enough to outweigh the energy needed to involve the s-
electrons. Hence, for aluminium and gallium +3 is the most stable state, but +1 gains importance for indium
and by thallium it becomes more stable than the +3 state. Nihonium is expected to continue this trend and
have +1 as its most stable oxidation state.[1]

The simplest possible nihonium compound is the monohydride, NhH. The bonding is provided by the 7p1/2
electron of nihonium and the 1s electron of hydrogen. The SO interaction causes the binding energy of
nihonium monohydride to be reduced by about 1 eV[1] and the nihonium–hydrogen bond length to
decrease as the bonding 7p1/2 orbital is relativistically contracted. This is unique among the 7p element
monohydrides; all the others have relativistic expansion of the bond length instead of contraction.[115]
Another effect of the SO interaction is that the Nh–H bond is expected to have significant pi bonding
character (side-on orbital overlap), unlike the almost pure sigma bonding (head-on orbital overlap) in
thallium monohydride (TlH).[116] The analogous monofluoride (NhF) should also exist.[112] Nihonium(I) is
predicted to be more similar to silver(I) than thallium(I):[1] the Nh+ ion is expected to more willingly bind
anions, so that NhCl should be quite soluble in excess hydrochloric acid or ammonia; TlCl is not. In
contrast to Tl+, which forms the strongly basic hydroxide (TlOH) in solution, the Nh+ cation should instead
hydrolyse all the way to the amphoteric oxide Nh2 O, which would be soluble in aqueous ammonia and
weakly soluble in water.[4]

The adsorption behaviour of nihonium on gold surfaces in thermochromatographical experiments is


expected to be closer to that of astatine than that of thallium. The destabilisation of the 7p3/2 subshell
effectively leads to a valence shell closing at the 7s2 7p2 configuration rather than the expected 7s2 7p6
configuration with its stable octet. As such, nihonium, like astatine, can be considered to be one p-electron
short of a closed valence shell. Hence, even though nihonium is in group 13, it has several properties similar
to the group 17 elements. (Tennessine in group 17 has some group-13-like properties, as it has three valence
electrons outside the 7s2 7p2 closed shell.[117]) Nihonium is expected to be able to gain an electron to attain
this closed-shell configuration, forming the −1 oxidation state like the halogens (fluorine, chlorine, bromine,
iodine, and astatine). This state should be more stable than it is for thallium as the SO splitting of the 7p
subshell is greater than that for the 6p subshell.[5] Nihonium should be the most electronegative of the
metallic group 13 elements,[1] even more electronegative than tennessine, the period 7 congener of the
halogens: in the compound NhTs, the negative charge is expected to be on the nihonium atom rather than
the tennessine atom.[112] The −1 oxidation should be more stable for nihonium than for tennessine.[1][118]
The electron affinity of nihonium is calculated to be around 0.68 eV, higher than thallium's at 0.4 eV;
tennessine's is expected to be 1.8 eV, the lowest in its group.[1] It is theoretically predicted that nihonium
should have an enthalpy of sublimation around 150 kJ/mol and an enthalpy of adsorption on a gold surface
around −159 kJ/mol.[119]

Significant 6d involvement is expected in the


Nh–Au bond, although it is expected to be more
unstable than the Tl–Au bond and entirely due to
magnetic interactions. This raises the possibility
of some transition metal character for
nihonium.[114] On the basis of the small energy
gap between the 6d and 7s electrons, the higher
oxidation states +3 and +5 have been suggested BCl3 has a trigonal NhCl3 is predicted to be T-
for nihonium. [1][4] Some simple compounds with structure. shaped.

nihonium in the +3 oxidation state would be the


trihydride (NhH3 ), trifluoride (NhF3 ), and
trichloride (NhCl3 ). These molecules are predicted to be T-shaped and not trigonal planar as their boron
analogues are:[q] this is due to the influence of the 6d5/2 electrons on the bonding.[116][r] The heavier
nihonium tribromide (NhBr3 ) and triiodide (NhI3 ) are trigonal planar due to the increased steric repulsion
between the peripheral atoms; accordingly, they do not show significant 6d involvement in their bonding,
though the large 7s–7p energy gap means that they show reduced sp2 hybridisation compared to their boron
analogues.[116]

The bonding in the lighter NhX3 molecules can be considered as that of a linear NhX+2 species (similar to

HgF2 or AuF2 ) with an additional Nh–X bond involving the 7p orbital of nihonium perpendicular to the
other two ligands. These compounds are all expected to be highly unstable towards the loss of an X2
molecule and reduction to nihonium(I):[116]

NhX3 → NhX + X2

Nihonium thus continues the trend down group 13 of reduced stability of the +3 oxidation state, as all five
of these compounds have lower reaction energies than the unknown thallium(III) iodide.[s] The +3 state is
stabilised for thallium in anionic complexes such as TlI−4, and the presence of a possible vacant coordination
site on the lighter T-shaped nihonium trihalides is expected to allow a similar stabilisation of NhF−4 and
perhaps NhCl−4.[116]

The +5 oxidation state is unknown for all lighter group 13 elements: calculations predict that nihonium
pentahydride (NhH5 ) and pentafluoride (NhF5 ) should have a square pyramidal molecular geometry, but
also that both would be highly thermodynamically unstable to loss of an X2 molecule and reduction to
nihonium(III). Again, some stabilisation is expected for anionic complexes, such as NhF−6. The structures of
the nihonium trifluoride and pentafluoride molecules are the same as those for chlorine trifluoride and
pentafluoride.[116]

Experimental chemistry
The chemical characteristics of nihonium have yet to be determined unambiguously.[119][124] The isotopes
284 Nh, 285 Nh, and 286 Nh have half-lives long enough for chemical investigation.[119] From 2010 to 2012,

some preliminary chemical experiments were performed at the JINR to determine the volatility of nihonium.
The isotope 284 Nh was investigated, made as the daughter of 288 Mc produced in the 243 Am+48 Ca
reaction. The nihonium atoms were synthesised in a recoil chamber and then carried along
polytetrafluoroethylene (PTFE) capillaries at 70 °C by a carrier gas to the gold-covered detectors. About ten
to twenty atoms of 284 Nh were produced, but none of these atoms were registered by the detectors,
suggesting either that nihonium was similar in volatility to the noble gases (and thus diffused away too
quickly to be detected) or, more plausibly, that pure nihonium was not very volatile and thus could not
efficiently pass through the PTFE capillaries.[119] Formation of the hydroxide NhOH should ease the
transport, as nihonium hydroxide is expected to be more volatile than elemental nihonium, and this reaction
could be facilitated by adding more water vapour into the carrier gas. It seems likely that this formation is
not kinetically favoured, so the longer-lived isotopes 285 Nh and 286 Nh were considered more desirable for
future experiments.[119][125]

A 2017 experiment at the JINR, producing 284 Nh and 285 Nh via the 243 Am+48 Ca reaction as the
daughters of 288 Mc and 289 Mc, avoided this problem by removing the quartz surface, using only PTFE.
No nihonium atoms were observed after chemical separation, implying an unexpectedly large retention of
nihonium atoms on PTFE surfaces. This experimental result for the interaction limit of nihonium atoms with
PTFE
a PTFE surface (−ΔHads (Nh) > 45 kJ/mol) disagrees significantly with previous theory, which expected a
lower value of 14.00 kJ/mol. This suggests that the nihonium species involved in the previous experiment
was likely not elemental nihonium but rather nihonium hydroxide, and that high-temperature techniques
such as vacuum chromatography would be necessary to further probe the behaviour of elemental
nihonium.[126] Bromine saturated with boron tribromide has been suggested as a carrier gas for experiments
on nihonium chemistry; this oxidises nihonium's lighter congener thallium to thallium(III), providing an
avenue to investigate the oxidation states of nihonium, similar to earlier experiments done on the bromides
of group 5 elements, including the superheavy dubnium.[127]

See also
Chemistry portal

Notes
a. 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[14] or 112;[15] sometimes, the term is presented an
equivalent to the term "transactinide", which puts an upper limit before the beginning of the
hypothetical superactinide series).[16] 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.
b. 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.[17] In comparison, the reaction that resulted in hassium discovery, 208Pb
+19
+ 58Fe, had a cross section of ~20 pb (more specifically, 19-11 pb), as estimated by the
discoverers.[18]
c. 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 1 28 1
14Si + 0n → 13Al + 1p 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.[22]
d. This figure also marks the generally accepted upper limit for lifetime of a compound
nucleus.[27]
e. 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.[29] 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.[30]
f. Not all decay modes are caused by electrostatic repulsion. For example, beta decay is
caused by the weak interaction.[37]
g. 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.[42]
h. 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.[47] The first direct
measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[48] 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).[49]
i. 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).[38] 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.
j. Spontaneous fission was discovered by Soviet physicist Georgy Flerov,[50] a leading
scientist at JINR, and thus it was a "hobbyhorse" for the facility.[51] 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.[27] They thus preferred to link new isotopes to the already known ones by
successive alpha decays.[50]
k. For instance, element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics
in Stockholm, Stockholm County, Sweden.[52] 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.[53]
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.[53] JINR insisted that they
were the first to create the element and suggested a name of their own for the new element,
joliotium;[54] the Soviet name was also not accepted (JINR later referred to the naming of the
element 102 as "hasty").[55] This name was proposed to IUPAC in a written response to their
ruling on priority of discovery claims of elements, signed 29 September 1992.[55] The name
"nobelium" remained unchanged on account of its widespread usage.[56]
l. Transactinide elements, such as nihonium, are produced by nuclear fusion. These fusion
reactions can be divided into "hot" and "cold" fusion, depending on the excitation energy of
the compound nucleus produced. "Cold fusion" in the context of superheavy element
synthesis is a distinct concept from the idea that nuclear fusion can be achieved under room
temperature conditions.[57] In hot fusion reactions, light, high-energy projectiles are
accelerated towards heavy targets (actinides), creating compound nuclei at high excitation
energy (~40–50 MeV) that may fission, or alternatively emit several (3 to 5) neutrons.[58] Cold
fusion reactions use heavier projectiles, typically from the fourth period, and lighter targets,
usually lead and bismuth. The fused nuclei produced have a relatively low excitation energy
(~10–20 MeV), which decreases the probability that they will undergo fission reactions. As
the fused nuclei cool to the ground state, they emit only one or two neutrons. Hot fusion
produces more neutron-rich products because actinides have the highest neutron-to-proton
ratios of any elements, and is currently the only method to produce the superheavy elements
from flerovium (element 114) onwards.[59]
m. Neptunium had been first reported at Riken by Nishina and Kenjiro Kimura in 1940, who did
not get naming rights because they could not chemically separate and identify their
discovery.[97][98]
n. Different sources give different values for half-lives; the most recently published values are
listed.
o. This isotope is unconfirmed
p. The quantum number corresponds to the letter in the electron orbital name: 0 to s, 1 to p, 2 to
d, etc.
q. Among the stable group 13 elements, only boron forms monomeric halides at standard
conditions; those of aluminium, gallium, indium, and thallium form ionic lattice structures or
(in a few cases) dimerise.[120][121]
r. The opposite effect is expected for the superheavy member of group 17, tennessine, due to
the relativistic stabilisation of the 7p1/2 orbital: thus IF3 is T-shaped, but TsF3 is expected to
be trigonal planar.[122]
s. The compound with stoichiometry TlI3 is a thallium(I) compound involving the triiodide anion,
I−3 .[123]

References
1. Hoffman, Darleane C.; Lee, Diana M.; Pershina, Valeria (2006). "Transactinides and the
future elements". In Morss; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the
Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer
Science+Business Media. ISBN 978-1-4020-3555-5.
2. Seaborg, Glenn T. (c. 2006). "transuranium element (chemical element)" (http://www.britanni
ca.com/EBchecked/topic/603220/transuranium-element). Encyclopædia Britannica.
Retrieved 16 March 2010.
3. Bonchev, Danail; Kamenska, Verginia (1981). "Predicting the Properties of the 113–120
Transactinide Elements" (https://www.researchgate.net/publication/239657207_Predicting_t
he_properties_of_the_113_to_120_transactinide_elements). Journal of Physical Chemistry.
85 (9): 1177–1186. doi:10.1021/j150609a021 (https://doi.org/10.1021%2Fj150609a021).
4. Fricke, Burkhard (1975). "Superheavy elements: a prediction of their chemical and physical
properties" (https://www.researchgate.net/publication/225672062). Recent Impact of Physics
on Inorganic Chemistry. Structure and Bonding. 21: 89–144. doi:10.1007/BFb0116498 (http
s://doi.org/10.1007%2FBFb0116498). ISBN 978-3-540-07109-9. Retrieved 4 October 2013.
5. Thayer, John S. (2010). "Relativistic Effects and the Chemistry of the Heavier Main Group
Elements". In Barysz, Maria; Ishikawa, Yasuyuki (eds.). Relativistic Methods for Chemists.
Challenges and Advances in Computational Chemistry and Physics. Vol. 10. Springer.
pp. 63–67. doi:10.1007/978-1-4020-9975-5_2 (https://doi.org/10.1007%2F978-1-4020-9975-
5_2). ISBN 978-1-4020-9974-8.
6. Keller, O. L. Jr.; Burnett, J. L.; Carlson, T. A.; Nestor, C. W. Jr. (1969). "Predicted Properties of
the Super Heavy Elements. I. Elements 113 and 114, Eka-Thallium and Eka-Lead". The
Journal of Physical Chemistry. 74 (5): 1127−1134. doi:10.1021/j100700a029 (https://doi.org/
10.1021%2Fj100700a029).
7. Atarah, Samuel A.; Egblewogbe, Martin N. H.; Hagoss, Gebreyesus G. (2020). "First
principle study of the structural and electronic properties of Nihonium". MRS Advances: 1–9.
doi:10.1557/adv.2020.159 (https://doi.org/10.1557%2Fadv.2020.159).
8. Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020
evaluation of nuclear properties" (https://www-nds.iaea.org/amdc/ame2020/NUBASE2020.p
df) (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae (https://doi.or
g/10.1088%2F1674-1137%2Fabddae).
9. Hofmann, S.; Heinz, S.; Mann, R.; Maurer, J.; Münzenberg, G.; Antalic, S.; Barth, W.; et al.
(2016). "Remarks on the Fission Barriers of SHN and Search for Element 120". In
Peninozhkevich, Yu. E.; Sobolev, Yu. G. (eds.). Exotic Nuclei: EXON-2016 Proceedings of
the International Symposium on Exotic Nuclei. Exotic Nuclei. pp. 155–164.
ISBN 9789813226555.
10. Hofmann, S.; Heinz, S.; Mann, R.; Maurer, J.; Münzenberg, G.; Antalic, S.; Barth, W.; et al.
(2016). "Review of even element super-heavy nuclei and search for element 120". The
European Physics Journal A. 2016 (52). doi:10.1140/epja/i2016-16180-4 (https://doi.org/10.1
140%2Fepja%2Fi2016-16180-4).
11. "WebElements Periodic Table » Nihonium » the essentials" (https://www.webelements.com/
nihonium/). www.webelements.com. Retrieved 8 June 2024.
12. "Nihonium | Nh (Element) - PubChem" (https://pubchem.ncbi.nlm.nih.gov/element/Nihonium
#section=History). pubchem.ncbi.nlm.nih.gov. Retrieved 8 June 2024.
13. "Nihonium (Nh) | AMERICAN ELEMENTS ®" (https://www.americanelements.com/nihoniu
m.html). American Elements: The Materials Science Company. Retrieved 24 April 2024.
14. Krämer, K. (2016). "Explainer: superheavy elements" (https://www.chemistryworld.com/news/
explainer-superheavy-elements/1010345.article). Chemistry World. Retrieved 15 March
2020.
15. "Discovery of Elements 113 and 115" (https://web.archive.org/web/20150911081623/https://
pls.llnl.gov/research-and-development/nuclear-science/project-highlights/livermorium/eleme
nts-113-and-115). Lawrence Livermore National Laboratory. Archived from the original (http
s://pls.llnl.gov/research-and-development/nuclear-science/project-highlights/livermorium/ele
ments-113-and-115) on 11 September 2015. Retrieved 15 March 2020.
16. 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 (https://doi.org/10.1002%2F
9781119951438.eibc2632). ISBN 978-1-119-95143-8. S2CID 127060181 (https://api.semant
icscholar.org/CorpusID:127060181).
17. 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 (https://doi.org/10.1103%2FPhysRevC.79.0246
08). ISSN 0556-2813 (https://www.worldcat.org/issn/0556-2813).
18. Münzenberg, G.; Armbruster, P.; Folger, H.; et al. (1984). "The identification of element 108"
(https://web.archive.org/web/20150607124040/http://www.gsi-heavy-ion-researchcenter.org/f
orschung/kp/kp2/ship/108-discovery.pdf) (PDF). Zeitschrift für Physik A. 317 (2): 235–236.
Bibcode:1984ZPhyA.317..235M (https://ui.adsabs.harvard.edu/abs/1984ZPhyA.317..235M).
doi:10.1007/BF01421260 (https://doi.org/10.1007%2FBF01421260). S2CID 123288075 (htt
ps://api.semanticscholar.org/CorpusID:123288075). Archived from the original (http://www.gs
i-heavy-ion-researchcenter.org/forschung/kp/kp2/ship/108-discovery.pdf) (PDF) on 7 June
2015. Retrieved 20 October 2012.
19. Subramanian, S. (28 August 2019). "Making New Elements Doesn't Pay. Just Ask This
Berkeley Scientist" (https://www.bloomberg.com/news/features/2019-08-28/making-new-ele
ments-doesn-t-pay-just-ask-this-berkeley-scientist). Bloomberg Businessweek. Retrieved
18 January 2020.
20. Ivanov, D. (2019). "Сверхтяжелые шаги в неизвестное" (https://nplus1.ru/material/2019/03/
25/120-element) [Superheavy steps into the unknown]. nplus1.ru (in Russian). Retrieved
2 February 2020.
21. Hinde, D. (2017). "Something new and superheavy at the periodic table" (https://theconversa
tion.com/something-new-and-superheavy-at-the-periodic-table-26286). The Conversation.
Retrieved 30 January 2020.
22. 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 (https://ui.adsa
bs.harvard.edu/abs/1959NucPh..10..226K). doi:10.1016/0029-5582(59)90211-1 (https://doi.o
rg/10.1016%2F0029-5582%2859%2990211-1).
23. 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" (https://doi.org/10.1051%2Fepjconf%2F20158600061). European Physical
Journal Web of Conferences. 86: 00061. Bibcode:2015EPJWC..8600061W (https://ui.adsab
s.harvard.edu/abs/2015EPJWC..8600061W). doi:10.1051/epjconf/20158600061 (https://doi.
org/10.1051%2Fepjconf%2F20158600061). hdl:1885/148847 (https://hdl.handle.net/1885%
2F148847). ISSN 2100-014X (https://www.worldcat.org/issn/2100-014X).
24. "Nuclear Reactions" (http://oregonstate.edu/instruct/ch374/ch418518/Chapter%2010%20NU
CLEAR%20REACTIONS.pdf) (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 (htt
ps://doi.org/10.1002%2F0471768626.ch10). ISBN 978-0-471-76862-3.
25. Krása, A. (2010). "Neutron Sources for ADS". Faculty of Nuclear Sciences and Physical
Engineering. Czech Technical University in Prague: 4–8. S2CID 28796927 (https://api.sema
nticscholar.org/CorpusID:28796927).
26. Wapstra, A. H. (1991). "Criteria that must be satisfied for the discovery of a new chemical
element to be recognized" (http://publications.iupac.org/pac/pdf/1991/pdf/6306x0879.pdf)
(PDF). Pure and Applied Chemistry. 63 (6): 883. doi:10.1351/pac199163060879 (https://doi.
org/10.1351%2Fpac199163060879). ISSN 1365-3075 (https://www.worldcat.org/issn/1365-3
075). S2CID 95737691 (https://api.semanticscholar.org/CorpusID:95737691).
27. Hyde, E. K.; Hoffman, D. C.; Keller, O. L. (1987). "A History and Analysis of the Discovery of
Elements 104 and 105" (http://www.escholarship.org/uc/item/05x8w9h7). Radiochimica Acta.
42 (2): 67–68. doi:10.1524/ract.1987.42.2.57 (https://doi.org/10.1524%2Fract.1987.42.2.57).
ISSN 2193-3405 (https://www.worldcat.org/issn/2193-3405). S2CID 99193729 (https://api.se
manticscholar.org/CorpusID:99193729).
28. Chemistry World (2016). "How to Make Superheavy Elements and Finish the Periodic Table
[Video]" (https://www.scientificamerican.com/article/how-to-make-superheavy-elements-and-
finish-the-periodic-table-video/). Scientific American. Retrieved 27 January 2020.
29. Hoffman, Ghiorso & Seaborg 2000, p. 334.
30. Hoffman, Ghiorso & Seaborg 2000, p. 335.
31. Zagrebaev, Karpov & Greiner 2013, p. 3.
32. Beiser 2003, p. 432.
33. Pauli, N. (2019). "Alpha decay" (http://metronu.ulb.ac.be/npauly/Pauly/physnu/chapter_5.pdf)
(PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part).
Université libre de Bruxelles. Retrieved 16 February 2020.
34. Pauli, N. (2019). "Nuclear fission" (http://metronu.ulb.ac.be/npauly/Pauly/physnu/chapter_8.p
df) (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part).
Université libre de Bruxelles. Retrieved 16 February 2020.
35. Staszczak, A.; Baran, A.; Nazarewicz, W. (2013). "Spontaneous fission modes and lifetimes
of superheavy elements in the nuclear density functional theory" (https://doi.org/10.1103%2F
physrevc.87.024320). Physical Review C. 87 (2): 024320–1. arXiv:1208.1215 (https://arxiv.or
g/abs/1208.1215). Bibcode:2013PhRvC..87b4320S (https://ui.adsabs.harvard.edu/abs/2013
PhRvC..87b4320S). doi:10.1103/physrevc.87.024320 (https://doi.org/10.1103%2Fphysrevc.
87.024320). ISSN 0556-2813 (https://www.worldcat.org/issn/0556-2813).
36. Audi et al. 2017, pp. 030001-129–030001-138.
37. Beiser 2003, p. 439.
38. Beiser 2003, p. 433.
39. Audi et al. 2017, p. 030001-125.
40. 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
(https://ui.adsabs.harvard.edu/abs/2017EPJA...53..158A). doi:10.1140/epja/i2017-12348-8
(https://doi.org/10.1140%2Fepja%2Fi2017-12348-8). ISSN 1434-6001 (https://www.worldcat.
org/issn/1434-6001). S2CID 125849923 (https://api.semanticscholar.org/CorpusID:1258499
23).
41. Beiser 2003, p. 432–433.
42. Oganessian, Yu. (2012). "Nuclei in the "Island of Stability" of Superheavy Elements" (https://
doi.org/10.1088%2F1742-6596%2F337%2F1%2F012005). Journal of Physics: Conference
Series. 337 (1): 012005-1–012005-6. Bibcode:2012JPhCS.337a2005O (https://ui.adsabs.ha
rvard.edu/abs/2012JPhCS.337a2005O). doi:10.1088/1742-6596/337/1/012005 (https://doi.or
g/10.1088%2F1742-6596%2F337%2F1%2F012005). ISSN 1742-6596 (https://www.worldca
t.org/issn/1742-6596).
43. Moller, P.; Nix, J. R. (1994). Fission properties of the heaviest elements (https://digital.library.
unt.edu/ark:/67531/metadc674703/m2/1/high_res_d/32502.pdf) (PDF). Dai 2 Kai Hadoron
Tataikei no Simulation Symposium, Tokai-mura, Ibaraki, Japan. University of North Texas.
Retrieved 16 February 2020.
44. Oganessian, Yu. Ts. (2004). "Superheavy elements" (https://physicsworld.com/a/superheavy-
elements/). Physics World. 17 (7): 25–29. doi:10.1088/2058-7058/17/7/31 (https://doi.org/10.
1088%2F2058-7058%2F17%2F7%2F31). Retrieved 16 February 2020.
45. Schädel, M. (2015). "Chemistry of the superheavy elements" (https://doi.org/10.1098%2Frst
a.2014.0191). Philosophical Transactions of the Royal Society A: Mathematical, Physical
and Engineering Sciences. 373 (2037): 20140191. Bibcode:2015RSPTA.37340191S (http
s://ui.adsabs.harvard.edu/abs/2015RSPTA.37340191S). doi:10.1098/rsta.2014.0191 (https://
doi.org/10.1098%2Frsta.2014.0191). ISSN 1364-503X (https://www.worldcat.org/issn/1364-5
03X). PMID 25666065 (https://pubmed.ncbi.nlm.nih.gov/25666065).
46. Hulet, E. K. (1989). Biomodal spontaneous fission. 50th Anniversary of Nuclear Fission,
Leningrad, USSR. Bibcode:1989nufi.rept...16H (https://ui.adsabs.harvard.edu/abs/1989nufi.r
ept...16H).
47. Oganessian, Yu. Ts.; Rykaczewski, K. P. (2015). "A beachhead on the island of stability" (http
s://doi.org/10.1063%2FPT.3.2880). Physics Today. 68 (8): 32–38.
Bibcode:2015PhT....68h..32O (https://ui.adsabs.harvard.edu/abs/2015PhT....68h..32O).
doi:10.1063/PT.3.2880 (https://doi.org/10.1063%2FPT.3.2880). ISSN 0031-9228 (https://ww
w.worldcat.org/issn/0031-9228). OSTI 1337838 (https://www.osti.gov/biblio/1337838).
S2CID 119531411 (https://api.semanticscholar.org/CorpusID:119531411).
48. Grant, A. (2018). "Weighing the heaviest elements". Physics Today.
doi:10.1063/PT.6.1.20181113a (https://doi.org/10.1063%2FPT.6.1.20181113a).
S2CID 239775403 (https://api.semanticscholar.org/CorpusID:239775403).
49. Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table" (http
s://cen.acs.org/physical-chemistry/periodic-table/IYPT-Exploring-the-superheavy-elements-a
t-the-end-of-the-periodic-table/97/i21). Chemical & Engineering News. Retrieved 27 January
2020.
50. Robinson, A. E. (2019). "The Transfermium Wars: Scientific Brawling and Name-Calling
during the Cold War" (https://www.sciencehistory.org/distillations/the-transfermium-wars-scie
ntific-brawling-and-name-calling-during-the-cold-war). Distillations. Retrieved 22 February
2020.
51. "Популярная библиотека химических элементов. Сиборгий (экавольфрам)" (http://n-t.ru/
ri/ps/pb106.htm) [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.
52. "Nobelium - Element information, properties and uses | Periodic Table" (https://www.rsc.org/p
eriodic-table/element/102/nobelium). Royal Society of Chemistry. Retrieved 1 March 2020.
53. Kragh 2018, pp. 38–39.
54. Kragh 2018, p. 40.
55. 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" (https://www.iupac.org/publications/pac/1993/pdf/6508x1815.
pdf) (PDF). Pure and Applied Chemistry. 65 (8): 1815–1824. doi:10.1351/pac199365081815
(https://doi.org/10.1351%2Fpac199365081815). S2CID 95069384 (https://api.semanticschol
ar.org/CorpusID:95069384). Archived (https://web.archive.org/web/20131125223512/http://w
ww.iupac.org/publications/pac/1993/pdf/6508x1815.pdf) (PDF) from the original on 25
November 2013. Retrieved 7 September 2016.
56. Commission on Nomenclature of Inorganic Chemistry (1997). "Names and symbols of
transfermium elements (IUPAC Recommendations 1997)" (http://publications.iupac.org/pac/
pdf/1997/pdf/6912x2471.pdf) (PDF). Pure and Applied Chemistry. 69 (12): 2471–2474.
doi:10.1351/pac199769122471 (https://doi.org/10.1351%2Fpac199769122471).
57. Fleischmann, Martin; Pons, Stanley (1989). "Electrochemically induced nuclear fusion of
deuterium". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 261 (2):
301–308. doi:10.1016/0022-0728(89)80006-3 (https://doi.org/10.1016%2F0022-0728%288
9%2980006-3).
58. Barber, Robert C.; Gäggeler, Heinz W.; Karol, Paul J.; Nakahara, Hiromichi; Vardaci,
Emanuele; Vogt, Erich (2009). "Discovery of the element with atomic number 112 (IUPAC
Technical Report)" (https://doi.org/10.1351%2FPAC-REP-08-03-05). Pure and Applied
Chemistry. 81 (7): 1331. doi:10.1351/PAC-REP-08-03-05 (https://doi.org/10.1351%2FPAC-R
EP-08-03-05).
59. Armbruster, Peter; Munzenberg, Gottfried (1989). "Creating superheavy elements". Scientific
American. 34: 36–42.
60. Chapman, Kit (30 November 2016). "What it takes to make a new element" (https://www.che
mistryworld.com/what-it-takes-to-make-a-new-element/1017677.article). Chemistry World.
Royal Society of Chemistry. Retrieved 3 December 2016.
61. Hofmann, Sigurd (2016). The discovery of elements 107 to 112 (https://www.epj-conference
s.org/articles/epjconf/pdf/2016/26/epjconf-NS160-06001.pdf) (PDF). Nobel Symposium
NS160 – Chemistry and Physics of Heavy and Superheavy Elements.
doi:10.1051/epjconf/201613106001 (https://doi.org/10.1051%2Fepjconf%2F201613106001).
62. Oganessian, Yu. Ts.; et al. (1999). "Synthesis of Superheavy Nuclei in the 48Ca + 244Pu
Reaction" (https://web.archive.org/web/20200730232521/http://flerovlab.jinr.ru/linkc/flnr_pres
entations/articles/synthesis_of_Element_114_1999.pdf) (PDF). Physical Review Letters. 83
(16): 3154. Bibcode:1999PhRvL..83.3154O (https://ui.adsabs.harvard.edu/abs/1999PhRvL..
83.3154O). doi:10.1103/PhysRevLett.83.3154 (https://doi.org/10.1103%2FPhysRevLett.83.3
154). Archived from the original (http://flerovlab.jinr.ru/linkc/flnr_presentations/articles/synthes
is_of_Element_114_1999.pdf) (PDF) on 30 July 2020. Retrieved 5 April 2017.
63. Oganessian, Yu. Ts.; et al. (2004). "Measurements of cross sections and decay properties of
the isotopes of elements 112, 114, and 116 produced in the fusion reactions 233,238U,
242Pu, and 248Cm + 48Ca" (https://web.archive.org/web/20080528130343/http://www.jinr.r
u/publish/Preprints/2004/160%28E7-2004-160%29.pdf) (PDF). Physical Review C. 70 (6):
064609. Bibcode:2004PhRvC..70f4609O (https://ui.adsabs.harvard.edu/abs/2004PhRvC..70
f4609O). doi:10.1103/PhysRevC.70.064609 (https://doi.org/10.1103%2FPhysRevC.70.0646
09). Archived from the original (http://www.jinr.ru/publish/Preprints/2004/160(E7-2004-160).p
df) (PDF) on 28 May 2008.
64. Oganessian, Yu. Ts.; et al. (2000). "Synthesis of superheavy nuclei in the 48Ca + 244Pu
reaction: 288114". Physical Review C. 62 (4): 041604. Bibcode:2000PhRvC..62d1604O (htt
ps://ui.adsabs.harvard.edu/abs/2000PhRvC..62d1604O). doi:10.1103/PhysRevC.62.041604
(https://doi.org/10.1103%2FPhysRevC.62.041604).
65. Oganessian, Yu. Ts.; et al. (2004). "Measurements of cross sections for the fusion-
evaporation reactions 244Pu(48Ca,xn)292−x114 and 245Cm(48Ca,xn)293−x116" (http://lin
k.aps.org/abstract/PRC/V69/E054607/). Physical Review C. 69 (5): 054607.
Bibcode:2004PhRvC..69e4607O (https://ui.adsabs.harvard.edu/abs/2004PhRvC..69e4607
O). doi:10.1103/PhysRevC.69.054607 (https://doi.org/10.1103%2FPhysRevC.69.054607).
66. Oganessian, Yu. Ts.; Utyonkoy, V.; Lobanov, Yu.; Abdullin, F.; Polyakov, A.; Shirokovsky, I.;
Tsyganov, Yu.; Gulbekian, G.; Bogomolov, S.; Mezentsev, A. N.; et al. (2004). "Experiments
on the synthesis of element 115 in the reaction 243Am(48Ca,xn)291−x115" (https://web.arch
ive.org/web/20200307233820/http://lt-jds.jinr.ru/record/7440/files/178%28E7-2003-178%29.
pdf) (PDF). Physical Review C. 69 (2): 021601. Bibcode:2004PhRvC..69b1601O (https://ui.a
dsabs.harvard.edu/abs/2004PhRvC..69b1601O). doi:10.1103/PhysRevC.69.021601 (https://
doi.org/10.1103%2FPhysRevC.69.021601). Archived from the original (http://lt-jds.jinr.ru/reco
rd/7440/files/178%28E7-2003-178%29.pdf) (PDF) on 7 March 2020. Retrieved 13 December
2019.
67. Morita, Kōsuke (5 February 2016). "Q & A session" (https://www.youtube.com/watch?v=kGVk
kVMgvOg). The Foreign Correspondents' Club of Japan. Archived (https://ghostarchive.org/v
archive/youtube/20211114/kGVkkVMgvOg) from the original on 14 November 2021.
Retrieved 28 April 2017 – via YouTube.
68. Barber, Robert C.; Karol, Paul J; Nakahara, Hiromichi; Vardaci, Emanuele; Vogt, Erich W.
(2011). "Discovery of the elements with atomic numbers greater than or equal to 113 (IUPAC
Technical Report)" (https://doi.org/10.1351%2FPAC-REP-10-05-01). Pure Appl. Chem. 83
(7): 1485. doi:10.1351/PAC-REP-10-05-01 (https://doi.org/10.1351%2FPAC-REP-10-05-01).
69. Rudolph, D.; Forsberg, U.; Golubev, P.; Sarmiento, L. G.; Yakushev, A.; Andersson, L.-L.; Di
Nitto, A.; Düllmann, Ch. E.; Gates, J. M.; Gregorich, K. E.; Gross, C. J.; Heßberger, F. P.;
Herzberg, R.-D.; Khuyagbaatar, J.; Kratz, J. V.; Rykaczewski, K.; Schädel, M.; Åberg, S.;
Ackermann, D.; Block, M.; Brand, H.; Carlsson, B. G.; Cox, D.; Derkx, X.; Eberhardt, K.; Even,
J.; Fahlander, C.; Gerl, J.; Jäger, E.; Kindler, B.; Krier, J.; Kojouharov, I.; Kurz, N.; Lommel, B.;
Mistry, A.; Mokry, C.; Nitsche, H.; Omtvedt, J. P.; Papadakis, P.; Ragnarsson, I.; Runke, J.;
Schaffner, H.; Schausten, B.; Thörle-Pospiech, P.; Torres, T.; Traut, T.; Trautmann, N.; Türler,
A.; Ward, A.; Ward, D. E.; Wiehl, N. (2013). "Spectroscopy of Element 115 Decay Chains" (htt
p://lup.lub.lu.se/record/4002358). Physical Review Letters (Submitted manuscript). 111 (11):
112502. Bibcode:2013PhRvL.111k2502R (https://ui.adsabs.harvard.edu/abs/2013PhRvL.11
1k2502R). doi:10.1103/PhysRevLett.111.112502 (https://doi.org/10.1103%2FPhysRevLett.1
11.112502). ISSN 0031-9007 (https://www.worldcat.org/issn/0031-9007). PMID 24074079 (h
ttps://pubmed.ncbi.nlm.nih.gov/24074079). S2CID 3838065 (https://api.semanticscholar.org/
CorpusID:3838065).
70. Morita, Kosuke; Morimoto, Kouji; Kaji, Daiya; Akiyama, Takahiro; Goto, Sin-ichi; Haba,
Hiromitsu; Ideguchi, Eiji; Kanungo, Rituparna; Katori, Kenji; Koura, Hiroyuki; Kudo, Hisaaki;
Ohnishi, Tetsuya; Ozawa, Akira; Suda, Toshimi; Sueki, Keisuke; Xu, HuShan; Yamaguchi,
Takayuki; Yoneda, Akira; Yoshida, Atsushi; Zhao, YuLiang (2004). "Experiment on the
Synthesis of Element 113 in the Reaction 209Bi(70Zn,n)278113" (https://doi.org/10.1143%2
FJPSJ.73.2593). Journal of the Physical Society of Japan. 73 (10): 2593–2596.
Bibcode:2004JPSJ...73.2593M (https://ui.adsabs.harvard.edu/abs/2004JPSJ...73.2593M).
doi:10.1143/JPSJ.73.2593 (https://doi.org/10.1143%2FJPSJ.73.2593).
71. 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 (IUPAC Technical Report)" (https://doi.org/10.1515%2Fpac-2015-0502). Pure
Appl. Chem. 88 (1–2): 139–153. doi:10.1515/pac-2015-0502 (https://doi.org/10.1515%2Fpac
-2015-0502).
72. Dmitriev, S. N.; Oganessyan, Yu. Ts.; Utyonkov, V. K.; Shishkin, S. V.; Yeremin, A. V.;
Lobanov, Yu. V.; Tsyganov, Yu. S.; Chepygin, V. I.; Sokol, E. A.; Vostokin, G. K.; Aksenov, N.
V.; Hussonnois, M.; Itkis, M. G.; Gäggeler, H. W.; Schumann, D.; Bruchertseifer, H.; Eichler,
R.; Shaughnessy, D. A.; Wilk, P. A.; Kenneally, J. M.; Stoyer, M. A.; Wild, J. F. (2005).
"Chemical identification of dubnium as a decay product of element 115 produced in the
reaction 48Ca+243Am". Mendeleev Communications. 15 (1): 1–4.
doi:10.1070/MC2005v015n01ABEH002077 (https://doi.org/10.1070%2FMC2005v015n01A
BEH002077). S2CID 98386272 (https://api.semanticscholar.org/CorpusID:98386272).
73. Oganessian, Yu. Ts.; Utyonkov, V.; Dmitriev, S.; Lobanov, Yu.; Itkis, M.; Polyakov, A.;
Tsyganov, Yu.; Mezentsev, A.; Yeremin, A.; Voinov, A. A.; et al. (2005). "Synthesis of elements
115 and 113 in the reaction 243Am + 48Ca" (https://www.dora.lib4ri.ch/psi/islandora/object/p
si%3A13194). Physical Review C. 72 (3): 034611. Bibcode:2005PhRvC..72c4611O (https://
ui.adsabs.harvard.edu/abs/2005PhRvC..72c4611O). doi:10.1103/PhysRevC.72.034611 (htt
ps://doi.org/10.1103%2FPhysRevC.72.034611).
74. Morimoto, Kouji (2016). "The discovery of element 113 at RIKEN" (http://www.physics.adelai
de.edu.au/cssm/workshops/inpc2016/talks/Morimoto_Mon_HallL_0930.pdf) (PDF). 26th
International Nuclear Physics Conference. Retrieved 14 May 2017.
75. Oganessian, Yu. Ts.; Utyonkov, V.; Lobanov, Yu.; Abdullin, F.; Polyakov, A.; Sagaidak, R.;
Shirokovsky, I.; Tsyganov, Yu.; Voinov, A.; Gulbekian, Gulbekian; et al. (2007). "Synthesis of
the isotope 282113 in the 237Np + 48Ca fusion reaction" (http://nrv.jinr.ru/pdf_file/PhysRevC
_76_011601.pdf) (PDF). Physical Review C. 76 (1): 011601(R).
Bibcode:2007PhRvC..76a1601O (https://ui.adsabs.harvard.edu/abs/2007PhRvC..76a1601
O). doi:10.1103/PhysRevC.76.011601 (https://doi.org/10.1103%2FPhysRevC.76.011601).
76. Morita, Kosuke; Morimoto, Kouji; Kaji, Daiya; Haba, Hiromitsu; Ozeki, Kazutaka; Kudou,
Yuki; Sato, Nozomi; Sumita, Takayuki; Yoneda, Akira; Ichikawa, Takatoshi; Fujimori,
Yasuyuki; Goto, Sin-ichi; Ideguchi, Eiji; Kasamatsu, Yoshitaka; Katori, Kenji; Komori, Yukiko;
Koura, Hiroyuki; Kudo, Hisaaki; Ooe, Kazuhiro; Ozawa, Akira; Tokanai, Fuyuki; Tsukada,
Kazuaki; Yamaguchi, Takayuki; Yoshida, Atsushi (25 May 2009). "Decay Properties of
266Bh and 262Db Produced in the 248Cm + 23Na Reaction". Journal of the Physical
Society of Japan. 78 (6): 064201–1–6. arXiv:0904.1093 (https://arxiv.org/abs/0904.1093).
Bibcode:2009JPSJ...78f4201M (https://ui.adsabs.harvard.edu/abs/2009JPSJ...78f4201M).
doi:10.1143/JPSJ.78.064201 (https://doi.org/10.1143%2FJPSJ.78.064201).
S2CID 16415500 (https://api.semanticscholar.org/CorpusID:16415500).
77. Morimoto, Kouji; Morita, K.; Kaji, D.; Haba, H.; Ozeki, K.; Kudou, Y.; Sato, N.; Sumita, T.;
Yoneda, A.; Ichikawa, T.; Fujimori, Y.; Goto, S.; Ideguchi, E.; Kasamatsu, Y.; Katori, K.;
Komori, Y.; Koura, H.; Kudo, H.; Ooe, K.; Ozawa, A.; Tokanai, F.; Tsukada, K.; Yamaguchi, T.;
Yoshida, A. (October 2009). "Production and Decay Properties of 266Bh and its daughter
nuclei by using the 248Cm(23Na,5n)266Bh Reaction" (https://web.archive.org/web/2017092
1193318/http://www.kernchemie.uni-mainz.de/downloads/che_7/presentations/morimoto.pd
f) (PDF). Archived from the original (http://www.kernchemie.uni-mainz.de/downloads/che_7/p
resentations/morimoto.pdf) (PDF) on 21 September 2017. Retrieved 28 April 2017 – via
University of Mainz.
78. Oganessian, Yuri Ts.; Abdullin, F. Sh.; Bailey, P. D.; Benker, D. E.; Bennett, M. E.; Dmitriev, S.
N.; Ezold, J. G.; Hamilton, J. H.; Henderson, R. A.; Itkis, M. G.; Lobanov, Yuri V.; Mezentsev,
A. N.; Moody, K. J.; Nelson, S. L.; Polyakov, A. N.; Porter, C. E.; Ramayya, A. V.; Riley, F. D.;
Roberto, J. B.; Ryabinin, M. A.; Rykaczewski, K. P.; Sagaidak, R. N.; Shaughnessy, D. A.;
Shirokovsky, I. V.; Stoyer, M. A.; Subbotin, V. G.; Sudowe, R.; Sukhov, A. M.; Tsyganov, Yu.
S.; Utyonkov, Vladimir K.; Voinov, A. A.; Vostokin, G. K.; Wilk, P. A. (9 April 2010). "Synthesis
of a New Element with Atomic Number Z=117" (https://www.researchgate.net/publication/44
610795). Physical Review Letters. 104 (14): 142502. Bibcode:2010PhRvL.104n2502O (http
s://ui.adsabs.harvard.edu/abs/2010PhRvL.104n2502O).
doi:10.1103/PhysRevLett.104.142502 (https://doi.org/10.1103%2FPhysRevLett.104.14250
2). PMID 20481935 (https://pubmed.ncbi.nlm.nih.gov/20481935).
79. K. Morita; Morimoto, Kouji; Kaji, Daiya; Haba, Hiromitsu; Ozeki, Kazutaka; Kudou, Yuki;
Sumita, Takayuki; Wakabayashi, Yasuo; Yoneda, Akira; Tanaka, Kengo; et al. (2012). "New
Results in the Production and Decay of an Isotope, 278113, of the 113th Element". Journal
of the Physical Society of Japan. 81 (10): 103201. arXiv:1209.6431 (https://arxiv.org/abs/120
9.6431). Bibcode:2012JPSJ...81j3201M (https://ui.adsabs.harvard.edu/abs/2012JPSJ...81j3
201M). doi:10.1143/JPSJ.81.103201 (https://doi.org/10.1143%2FJPSJ.81.103201).
S2CID 119217928 (https://api.semanticscholar.org/CorpusID:119217928).
80. Chapman, Kit (8 February 2018). "Nihonium" (https://www.chemistryworld.com/podcasts/nih
onium/3008633.article). Chemistry World. Royal Society of Chemistry. Retrieved 20 March
2018.
81. Morita, Kosuke (2015). "SHE Research at RIKEN/GARIS" (http://cyclotron.tamu.edu/she201
5/assets/pdfs/presentations/Morita_SHE_2015_TAMU.pdf) (PDF). Retrieved 4 September
2018 – via Texas A&M University Cyclotron Institute.
82. "Existence of new element confirmed" (http://www.lunduniversity.lu.se/article/existence-of-ne
w-element-confirmed). Lund University. 27 August 2013. Retrieved 10 April 2016.
83. Gates, J. M.; Gregorich, K. E.; Gothe, O. .R; Uribe, E. C.; Pang, G. K.; Bleuel, D. L.; Block, M.;
Clark, R. M.; Campbell, C. M.; Crawford, H. L.; Cromaz, M.; Di Nitto, A.; Düllmann, Ch. E.;
Esker, N. E.; Fahlander, C.; Fallon, P.; Farjadi, R. M.; Forsberg, U.; Khuyagbaatar, J.;
Loveland, W.; MacChiavelli, A. O.; May, E. M.; Mudder, P. R.; Olive, D. T.; Rice, A. C.;
Rissanen, J.; Rudolph, D.; Sarmiento, L. G.; Shusterman, J. A.; et al. (2015). "Decay
spectroscopy of element 115 daughters: 280Rg→276Mt and 276Mt→Bh" (https://doi.org/10.
1103%2FPhysRevC.92.021301). Physical Review C. 92 (2): 021301.
Bibcode:2015PhRvC..92b1301G (https://ui.adsabs.harvard.edu/abs/2015PhRvC..92b1301
G). doi:10.1103/PhysRevC.92.021301 (https://doi.org/10.1103%2FPhysRevC.92.021301).
84. "Element 113: Ununtrium Reportedly Synthesised In Japan" (http://www.huffingtonpost.com/
2012/09/26/element-113-created-synthetically-japan_n_1916253.html). Huffington Post.
September 2012. Retrieved 22 April 2013.
85. McKellar, Bruce (22–23 October 2016). "President's report to the meeting of the IUPAP
Council and Commission Chairs" (https://web.archive.org/web/20201102094153/http://iupa
p.org/wp-content/uploads/2016/09/Presidents-report-20161013-2.pdf) (PDF). International
Union of Pure and Applied Physics. Archived from the original (http://iupap.org/wp-content/up
loads/2016/09/Presidents-report-20161013-2.pdf) (PDF) on 2 November 2020. Retrieved
14 January 2018.
86. "Discovery of the new chemical elements with numbers 113, 115, 117 and 118" (http://www.ji
nr.ru/posts/discovery-of-the-new-chemical-elements-with-numbers-113-115-117-and-118-2/).
Joint Institute for Nuclear Research. 6 January 2016. Retrieved 14 January 2018.
87. "Discovery and Assignment of Elements with Atomic Numbers 113, 115, 117 and 118" (http
s://web.archive.org/web/20151231074712/http://www.iupac.org/news/news-detail/article/disc
overy-and-assignment-of-elements-with-atomic-numbers-113-115-117-and-118.html).
IUPAC. 30 December 2015. Archived from the original (http://www.iupac.org/news/news-deta
il/article/discovery-and-assignment-of-elements-with-atomic-numbers-113-115-117-and-118.
html) on 31 December 2015. Retrieved 8 September 2018.
88. 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" (http://
portal.research.lu.se/portal/files/9762047/PhysLettB760_293_2016.pdf) (PDF). Physics
Letters B. 760 (2016): 293–296. Bibcode:2016PhLB..760..293F (https://ui.adsabs.harvard.ed
u/abs/2016PhLB..760..293F). doi:10.1016/j.physletb.2016.07.008 (https://doi.org/10.1016%2
Fj.physletb.2016.07.008). Retrieved 2 April 2016.
89. Forsberg, Ulrika; Fahlander, Claes; Rudolph, Dirk (2016). Congruence of decay chains of
elements 113, 115, and 117 (http://www.epj-conferences.org/articles/epjconf/pdf/2016/26/epj
conf-NS160-02003.pdf) (PDF). Nobel Symposium NS160 – Chemistry and Physics of Heavy
and Superheavy Elements. doi:10.1051/epjconf/201613102003 (https://doi.org/10.1051%2F
epjconf%2F201613102003).
90. 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" (https://doi.org/10.108
8%2F1361-6471%2Faa7293). Journal of Physics G: Nuclear and Particle Physics. 44
(75107): 075107. Bibcode:2017JPhG...44g5107Z (https://ui.adsabs.harvard.edu/abs/2017JP
hG...44g5107Z). doi:10.1088/1361-6471/aa7293 (https://doi.org/10.1088%2F1361-6471%2F
aa7293).
91. Chatt, J. (1979). "Recommendations for the Naming of Elements of Atomic Numbers Greater
than 100" (https://doi.org/10.1351%2Fpac197951020381). Pure Appl. Chem. 51 (2): 381–
384. doi:10.1351/pac197951020381 (https://doi.org/10.1351%2Fpac197951020381).
92. Noorden, Richard Van (27 September 2012). "Element 113 at Last?" (https://www.scientifica
merican.com/article/element-113-at-last/). Scientific American.
93.新元素 番、日本の発見確実に 合成に3回成功
113 (http://www.nikkei.com/article/DGXNASD
G2604F_W2A920C1CR8000/). Nihon Keizai Shimbun (in Japanese). 27 September 2012.
Retrieved 13 October 2012.
94. "Proposed name for 113th element a fulfilled wish for Japanese researchers" (https://mainich
i.jp/english/articles/20160609/p2a/00m/0na/010000c). The Mainichi. 9 June 2016. Retrieved
29 April 2018.
95. "Naming 113th element 'nihonium' a tribute to Japanese public support: researcher" (https://
mainichi.jp/english/articles/20160609/p2a/00m/0na/014000c). The Mainichi. 9 June 2016.
Retrieved 29 April 2018.
96. "IUPAC Is Naming The Four New Elements Nihonium, Moscovium, Tennessine, And
Oganesson" (http://iupac.org/iupac-is-naming-the-four-new-elements-nihonium-moscovium-t
ennessine-and-oganesson/). IUPAC. 8 June 2016. Retrieved 8 June 2016.
97. Ikeda, Nagao (25 July 2011). "The discoveries of uranium 237 and symmetric fission – From
the archival papers of Nishina and Kimura" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3
171289). Proceedings of the Japan Academy, Series B: Physical and Biological Sciences.
87 (7): 371–376. Bibcode:2011PJAB...87..371I (https://ui.adsabs.harvard.edu/abs/2011PJA
B...87..371I). doi:10.2183/pjab.87.371 (https://doi.org/10.2183%2Fpjab.87.371).
PMC 3171289 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3171289). PMID 21785255 (h
ttps://pubmed.ncbi.nlm.nih.gov/21785255).
98. En'yo, Hideto (26 May 2017). "Bikkuban kara 113-ban genso nihoniumu made, genso sōsei
no 138 oku-nen" ビックバンから 113番元素ニホニウムまで、元素創成の138億年 (htt
ps://web.archive.org/web/20180129004320/https://www.ssken.gr.jp/MAINSITE/event/2017/2
0170526-generalmeeting/lecture-01/SSKEN_generalmeeting2017_EnyoHideto_presentatio
n.pdf) [From the Big Bang to the 113th element nihonium: element creation of 13.8 billion
years] (PDF) (in Japanese). Archived from the original (https://www.ssken.gr.jp/MAINSITE/ev
ent/2017/20170526-generalmeeting/lecture-01/SSKEN_generalmeeting2017_EnyoHideto_
presentation.pdf) (PDF) on 29 January 2018. Retrieved 28 January 2018.
99. "Japan scientists plan to name atomic element 113 'Nihonium' " (https://web.archive.org/web/
20160609135534/http://mainichi.jp/english/articles/20160608/p2g/00m/0fp/060000c).
Mainichi Shimbun. 8 June 2016. Archived from the original (http://mainichi.jp/english/articles/
20160608/p2g/00m/0fp/060000c) on 9 June 2016. "Japanese scientists who discovered the
atomic element 113 plan to name it "Nihonium", sources close to the matter said
Wednesday."
ニホニウム」有力 日本初の新元素名称案、国際機関が9日公表
100. " " (http://www.sankei.com/lif
e/news/160608/lif1606080005-n1.html) [Nihonium the most probable]. The Sankei Shimbun
(in Japanese). 6 June 2016. "Rather than initially proposed Japanium which is derived from
Latin or French, Morita group leader seems to stick to his own language."
101. "IUPAC Announces the Names of the Elements 113, 115, 117, and 118" (https://iupac.org/iu
pac-announces-the-names-of-the-elements-113-115-117-and-118). IUPAC. 30 November
2016. Retrieved 30 November 2016.
102. "Naming ceremony held for new element 'nihonium' " (https://web.archive.org/web/20180128
132830/http://newsonjapan.com/html/newsdesk/article/119326.php). News on Japan. 15
March 2017. Archived from the original (http://newsonjapan.com/html/newsdesk/article/1193
26.php) on 28 January 2018. Retrieved 28 January 2018.
103. Audi, G.; Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S. (2017). "The NUBASE2016
evaluation of nuclear properties" (https://www-nds.iaea.org/amdc/ame2016/NUBASE2016.p
df) (PDF). Chinese Physics C. 41 (3): 030001. Bibcode:2017ChPhC..41c0001A (https://ui.ad
sabs.harvard.edu/abs/2017ChPhC..41c0001A). doi:10.1088/1674-1137/41/3/030001 (https://
doi.org/10.1088%2F1674-1137%2F41%2F3%2F030001).
104. Oganessian, Yu. Ts.; Utyonkov, V. K.; Kovrizhnykh, N. D.; et al. (2022). "New isotope 286Mc
produced in the 243Am+48Ca reaction" (https://doi.org/10.1103%2FPhysRevC.106.06430
6). Physical Review C. 106 (64306): 064306. Bibcode:2022PhRvC.106f4306O (https://ui.ads
abs.harvard.edu/abs/2022PhRvC.106f4306O). doi:10.1103/PhysRevC.106.064306 (https://d
oi.org/10.1103%2FPhysRevC.106.064306). S2CID 254435744 (https://api.semanticscholar.
org/CorpusID:254435744).
105. Oganessian, Y.T. (2015). "Super-heavy element research" (https://www.researchgate.net/pub
lication/273327193). Reports on Progress in Physics. 78 (3): 036301.
Bibcode:2015RPPh...78c6301O (https://ui.adsabs.harvard.edu/abs/2015RPPh...78c6301O).
doi:10.1088/0034-4885/78/3/036301 (https://doi.org/10.1088%2F0034-4885%2F78%2F3%2
F036301). PMID 25746203 (https://pubmed.ncbi.nlm.nih.gov/25746203). S2CID 37779526
(https://api.semanticscholar.org/CorpusID:37779526).
106. Sonzogni, Alejandro. "Interactive Chart of Nuclides" (https://web.archive.org/web/200708071
70127/http://www.nndc.bnl.gov/chart/reCenter.jsp?z=113&n=173). National Nuclear Data
Center: Brookhaven National Laboratory. Archived from the original (http://www.nndc.bnl.go
v/chart/reCenter.jsp?z=113&n=173) on 7 August 2007. Retrieved 6 June 2008.
107. Forsberg, Ulrika (September 2016). "Recoil-α-fission and recoil-α–α-fission events observed
in the reaction 48Ca + 243Am". Nuclear Physics A. 953: 117–138. arXiv:1502.03030 (https://
arxiv.org/abs/1502.03030). Bibcode:2016NuPhA.953..117F (https://ui.adsabs.harvard.edu/a
bs/2016NuPhA.953..117F). doi:10.1016/j.nuclphysa.2016.04.025 (https://doi.org/10.1016%2
Fj.nuclphysa.2016.04.025). S2CID 55598355 (https://api.semanticscholar.org/CorpusID:555
98355).
108. Considine, Douglas M.; Considine, Glenn D. (1994). Van Nostrand's Scientific Encyclopedia
(8th ed.). Wiley-Interscience. p. 623. ISBN 978-1-4757-6918-0.
109. 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 (https://ui.adsabs.harvard.edu/abs/2017PhyS...92b3003O).
doi:10.1088/1402-4896/aa53c1 (https://doi.org/10.1088%2F1402-4896%2Faa53c1).
S2CID 125713877 (https://api.semanticscholar.org/CorpusID:125713877).
110. Audi, Georges; Bersillon, Olivier; Blachot, Jean; Wapstra, Aaldert Hendrik (2003), "The
NUBASE evaluation of nuclear and decay properties" (https://hal.archives-ouvertes.fr/in2p3-
00020241/document), Nuclear Physics A, 729: 3–128, Bibcode:2003NuPhA.729....3A (http
s://ui.adsabs.harvard.edu/abs/2003NuPhA.729....3A), doi:10.1016/j.nuclphysa.2003.11.001
(https://doi.org/10.1016%2Fj.nuclphysa.2003.11.001)
111. Chapman, Kit (30 November 2016). "What it takes to make a new element" (https://www.che
mistryworld.com/features/what-it-takes-to-make-a-new-element/1017677.article). Chemistry
World. Retrieved 26 June 2024.
112. Stysziński, Jacek (2010). "Why do we Need Relativistic Computational Methods?".
Relativistic Methods for Chemists. Challenges and Advances in Computational Chemistry
and Physics. Vol. 10. pp. 139–146. doi:10.1007/978-1-4020-9975-5_3 (https://doi.org/10.100
7%2F978-1-4020-9975-5_3). ISBN 978-1-4020-9974-8.
113. Fægri Jr., Knut; Saue, Trond (2001). "Diatomic molecules between very heavy elements of
group 13 and group 17: A study of relativistic effects on bonding" (https://doi.org/10.1063%2F
1.1385366). The Journal of Chemical Physics. 115 (6): 2456.
Bibcode:2001JChPh.115.2456F (https://ui.adsabs.harvard.edu/abs/2001JChPh.115.2456F).
doi:10.1063/1.1385366 (https://doi.org/10.1063%2F1.1385366).
114. Zaitsevskii, A.; van Wüllen, C.; Rusakov, A.; Titov, A. (September 2007). "Relativistic DFT
and ab initio calculations on the seventh-row superheavy elements: E113 – E114" (http://tan
11.jinr.ru/pdf/07_Sep/S_3/04_Titov.pdf) (PDF). Retrieved 17 February 2018.
115. Han, Young-Kyu; Bae, Cheolbeom; Son, Sang-Kil; Lee, Yoon Sup (2000). "Spin–orbit effects
on the transactinide p-block element monohydrides MH (M=element 113–118)". Journal of
Chemical Physics. 112 (6): 2684. Bibcode:2000JChPh.112.2684H (https://ui.adsabs.harvar
d.edu/abs/2000JChPh.112.2684H). doi:10.1063/1.480842 (https://doi.org/10.1063%2F1.480
842). S2CID 9959620 (https://api.semanticscholar.org/CorpusID:9959620).
116. Seth, Michael; Schwerdtfeger, Peter; Fægri, Knut (1999). "The chemistry of superheavy
elements. III. Theoretical studies on element 113 compounds" (https://doi.org/10.1063%2F1.
480168). Journal of Chemical Physics. 111 (14): 6422–6433.
Bibcode:1999JChPh.111.6422S (https://ui.adsabs.harvard.edu/abs/1999JChPh.111.6422S).
doi:10.1063/1.480168 (https://doi.org/10.1063%2F1.480168). S2CID 41854842 (https://api.s
emanticscholar.org/CorpusID:41854842).
117. Demidov, Yu. A. (15 February 2017). "Quantum chemical modelling of electronic structure of
nihonium and astatine compounds" (http://www.jinr.ru/posts/25597/). Flerov Laboratory of
Nuclear Reactions. Retrieved 12 June 2017.
118. Nash, Clinton S.; Bursten, Bruce E. (1999). "Spin−Orbit Effects, VSEPR Theory, and the
Electronic Structures of Heavy and Superheavy Group IVA Hydrides and Group VIIIA
Tetrafluorides. A Partial Role Reversal for Elements 114 and 118". J. Phys. Chem. A. 103 (3):
402–410. Bibcode:1999JPCA..103..402N (https://ui.adsabs.harvard.edu/abs/1999JPCA..10
3..402N). doi:10.1021/jp982735k (https://doi.org/10.1021%2Fjp982735k). PMID 27676357 (h
ttps://pubmed.ncbi.nlm.nih.gov/27676357).
119. Eichler, Robert (2013). "First foot prints of chemistry on the shore of the Island of Superheavy
Elements". Journal of Physics: Conference Series. 420 (1): 012003. arXiv:1212.4292 (https://
arxiv.org/abs/1212.4292). Bibcode:2013JPhCS.420a2003E (https://ui.adsabs.harvard.edu/a
bs/2013JPhCS.420a2003E). doi:10.1088/1742-6596/420/1/012003 (https://doi.org/10.108
8%2F1742-6596%2F420%2F1%2F012003). S2CID 55653705 (https://api.semanticscholar.
org/CorpusID:55653705).
120. Greenwood, N. N.; Earnshaw, A. (1998). Chemistry of the Elements (https://books.google.co
m/books?id=EvTI-ouH3SsC&q=chemistry+of+the+elements) (2nd ed.). Butterworth-
Heinemann. pp. 195, 233–235, 237–240. ISBN 978-0-7506-3365-9.
121. Downs, A.J. (31 May 1993). Chemistry of Aluminium, Gallium, Indium and Thallium (https://b
ooks.google.com/books?id=v-04Kn758yIC). Springer Science & Business Media. pp. 128–
137. ISBN 978-0-7514-0103-5.
122. 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 (https://ui.adsabs.harvard.edu/abs/2003JPCA..
107..852B). doi:10.1021/jp026531m (https://doi.org/10.1021%2Fjp026531m).
123. Tebbe, K.-F.; Georgy, U. (December 1986). "Die Kristallstrukturen von Rubidiumtriiodid und
Thalliumtriiodid". Acta Crystallographica C. C42 (12): 1675–1678.
Bibcode:1986AcCrC..42.1675T (https://ui.adsabs.harvard.edu/abs/1986AcCrC..42.1675T).
doi:10.1107/S0108270186090972 (https://doi.org/10.1107%2FS0108270186090972).
124. Düllmann, Christoph E. (2012). "Superheavy elements at GSI: a broad research program
with element 114 in the focus of physics and chemistry". Radiochimica Acta. 100 (2): 67–74.
doi:10.1524/ract.2011.1842 (https://doi.org/10.1524%2Fract.2011.1842). S2CID 100778491
(https://api.semanticscholar.org/CorpusID:100778491).
125. 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. 24–28. ISBN 978-3-642-37466-1.
126. Aksenov, Nikolay V.; Steinegger, Patrick; Abdullin, Farid Sh.; Albin, Yury V.; Bozhikov,
Gospodin A.; Chepigin, Viktor I.; Eichler, Robert; Lebedev, Vyacheslav Ya.; Mamudarov,
Alexander Sh.; Malyshev, Oleg N.; Petrushkin, Oleg V.; Polyakov, Alexander N.; Popov, Yury
A.; Sabel'nikov, Alexey V.; Sagaidak, Roman N.; Shirokovsky, Igor V.; Shumeiko, Maksim V.;
Starodub, Gennadii Ya.; Tsyganov, Yuri S.; Utyonkov, Vladimir K.; Voinov, Alexey A.;
Vostokin, Grigory K.; Yeremin, Alexander; Dmitriev, Sergey N. (July 2017). "On the volatility
of nihonium (Nh, Z = 113)". The European Physical Journal A. 53 (158): 158.
Bibcode:2017EPJA...53..158A (https://ui.adsabs.harvard.edu/abs/2017EPJA...53..158A).
doi:10.1140/epja/i2017-12348-8 (https://doi.org/10.1140%2Fepja%2Fi2017-12348-8).
S2CID 125849923 (https://api.semanticscholar.org/CorpusID:125849923).
127. Tereshatov, E. E.; Boltoeva, M. Yu.; Folden III, C. M. (2015). "Resin Ion Exchange and Liquid-
Liquid Extraction of Indium and Thallium from Chloride Media". Solvent Extraction and Ion
Exchange. 33 (6): 607. doi:10.1080/07366299.2015.1080529 (https://doi.org/10.1080%2F07
366299.2015.1080529). S2CID 94078206 (https://api.semanticscholar.org/CorpusID:940782
06).
Bibliography
Audi, G.; Kondev, F. G.; Wang, M.; et al. (2017). "The NUBASE2016 evaluation of nuclear
properties". Chinese Physics C. 41 (3): 030001. Bibcode:2017ChPhC..41c0001A (https://ui.a
dsabs.harvard.edu/abs/2017ChPhC..41c0001A). doi:10.1088/1674-1137/41/3/030001 (http
s://doi.org/10.1088%2F1674-1137%2F41%2F3%2F030001).
Beiser, A. (2003). Concepts of modern physics (6th ed.). McGraw-Hill. ISBN 978-0-07-
244848-1. OCLC 48965418 (https://www.worldcat.org/oclc/48965418).
Hoffman, D. C.; Ghiorso, A.; Seaborg, G. T. (2000). The Transuranium People: The Inside
Story. World Scientific. ISBN 978-1-78-326244-1.
Kragh, H. (2018). From Transuranic to Superheavy Elements: A Story of Dispute and
Creation. Springer. ISBN 978-3-319-75813-8.
Zagrebaev, V.; Karpov, A.; Greiner, W. (2013). "Future of superheavy element research:
Which nuclei could be synthesized within the next few years?". Journal of Physics:
Conference Series. 420 (1): 012001. arXiv:1207.5700 (https://arxiv.org/abs/1207.5700).
Bibcode:2013JPhCS.420a2001Z (https://ui.adsabs.harvard.edu/abs/2013JPhCS.420a2001
Z). doi:10.1088/1742-6596/420/1/012001 (https://doi.org/10.1088%2F1742-6596%2F420%2
F1%2F012001). ISSN 1742-6588 (https://www.worldcat.org/issn/1742-6588).
S2CID 55434734 (https://api.semanticscholar.org/CorpusID:55434734).

External links
Nihonium (http://www.periodicvideos.com/videos/113.htm) at The Periodic Table of Videos
(University of Nottingham)
Uut and Uup Add Their Atomic Mass to Periodic Table (http://www.radiochemistry.org/periodi
ctable/elements/115.html)
Discovery of Elements 113 and 115 (https://web.archive.org/web/20050623012629/http://ww
w-cms.llnl.gov/e113_115/images.html)
Superheavy elements (https://physicsworld.com/a/superheavy-elements/)
WebElements.com: Nihonium (http://www.webelements.com/nihonium/)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Nihonium&oldid=1231680208"

You might also like