Sunteți pe pagina 1din 24

Nihonium

From Wikipedia, the free encyclopedia

Jump to navigationJump to search

General properties

Pronunciation /nɪˈhoʊniəm/ (ni-HOH-nee-əm)

Mass number 286 (most stable isotope)

Nihonium in the periodic table

H H
yd eli
ro u
g m
e
n

Lit
B BCNi
OFl
N
hi
er or
ar
tr
xy
ue
uyll obogor
o
miu nogein
n
m nene
n

SM Al
Si
PS
C
Ar
oa ulic
hulf
hl
g
di
g mi
oos
ur
or
o
un ni
npin
n
mes uhe
iu m or
m us

PC STi
VCMIr
CNi
CZi
GAr
SKr
ot
al cta
ahr
aock
onc
all
er
se
el
yp
as
ci ani
nonbel
piu
mni
eto
si
u nuami
galt
pm
acni
n
um dmdi
uaerniu
m i um num
um es m
me

RSt
Yt Zi
Ni
MTe
R PSi
CIn
Ti
ATe
X
uro
tri rc
ool
ch
ut
hall
lv
adi
nnti
llu
e
bi
nti
u obiyb
nhoaer
dum rin
di
um niudeti
edimi
m ouo
um um euni
uuny mn
m m nmum m
um
m

CB
La
Pr
Pr
ETe
HTLu
Ta
RIri
Pl
GM
TLBi
PR
aar
nt
as
our
rbi
ol
hu
tet
nt
he
di
ati
ol
er
hesol
a
es
iu
ha
eo
mop
umi
liu
iu
al
ni
undcu
all
amod
iu
m
nu
dy
et
iu
mummuumury
iu
dut
ni
o
mmmi
hi
m mm m m (e
m hun
uu le m
mm m
e
nt
)

Fr
RA
Pr
NABEi
MLa
DB
MDRCNi
Fl
MLi
O
acti
ot
ep
mer
ns
en
wr
ub
oh
eit
ar
oher
os
ve
g
nc
di
ni
ac
tu
eri
ke
tei
de
en
ni
riu
ne
mepo
ov
co
ra
iu
utin
ni
ci
liu
ni
le
ci
um
riu
st
nt
er
ni
iu
vi
mn
mm
iu
uumuvi
ummagni
u
m
uor
es
mmm mumdtieci
m
miu
so
muni um n
mum
m
cope
rniciu
m←
niho
nium

flero
vium

Atomic number 113


(Z)

Group group 13 (boron group)

Period period 7

Block p-block

Element unknown chemical properties, but


category probably a post-transition metal

Electron 14 10 2 1 [1]
[Rn] 5f 6d 7s 7p (predicted)
configuration

Electrons per shell 2, 8, 18, 32, 32, 18, 3 (predicted)

Physical properties

Phase at STP [1][2][3]


solid (predicted)
Melting point [1]
700 K (430 °C, 810 °F) (predicted)

Boiling point 1430 K (1130 °C, 2070 °F) (predicted)


[1][4]

Density (near r.t.) 3 [4]


16 g/cm (predicted)

Heat of fusion [3]


7.61 kJ/mol (extrapolated)

Heat of [2][4]
130 kJ/mol (predicted)
vaporization

Atomic properties

Oxidation states [1][4]


(−1), (+1), (+3), (+5) (predicted)
[5]

Ionization ● 1st: 704.9 kJ/mol (predicted)


energies [1]

● 2nd: 2240 kJ/mol (predicted)


[4]

● 3rd: 3020 kJ/mol (predicted)


[4]

● (more)

Atomic radius [1]


empirical: 170 pm (predicted)

Covalent radius [3]


172–180 pm (extrapolated)

Other properties

Crystal structure hexagonal close-packed (hcp)

[6]
(extrapolated)

CAS Number 54084-70-7


History

Naming After Japan (Nihon in Japanese)

Discovery Riken (Japan, first undisputed claim


2004)

JINR (Russia) and Livermore (US, first


announcement 2003)

Main isotopes of nihonium

I A H D P
s b a e r
o u l c o
t n f a d
o - y u
p d l c
e a i m t
n f o
c e d
e ( e
t
1

2
)

2 s 2 α 2

9 y 8
n s
0 6
?
N R
h g
[7

2 s 5 α 2

8 y . 8
n 5
7 3

N s R
h ? g
[8

2 s 9 α 2

8 y . 8
n 5
6 2

N s R
h g
2 s 4 α 2

8 y . 8
n 2
5 1

N s R
h g

α 2

0
2 0 R
8 . g
s 9
4 y 1
n E 2
N
s C 8
h
4

C
n

2 s 7 α 2

8 y 5 7
n
3 9
m
N s R
h g

2 s 7 α 2

8 y 3 7
n
2 8
m
N s R
h g

2 s 1 α 2

7 y . 7
n 4
8 4

N m R
h s g

● view
● talk
● edit
| references

Nihonium is a synthetic chemical element with symbol Nh and atomic number 113. It is extremely
radioactive; its most stable known isotope, nihonium-286, has a half-life of about 10 seconds. In the
periodic table, nihonium is a transactinide element in the p-block. It is a member of period 7 and group 13
(boron group).

Nihonium was first reported to have been created in 2003 by a Russian–American collaboration at the Joint
Institute for Nuclear Research (JINR) in Dubna, Russia, and in 2004 by a team of Japanese scientists at
Riken in Wakō, Japan. The confirmation of their claims in the ensuing years involved independent teams of
scientists working in the United States, Germany, Sweden, and China, as well as the original claimants in
Russia and Japan. In 2015, the IUPAC/IUPAP Joint Working Party recognised the element and assigned
the priority of the discovery and naming rights for the element to Riken, as it judged that they had
demonstrated that they had observed element 113 before the JINR team did so. The Riken team suggested
the name nihonium in 2016, which was approved in the same year. The name comes from the
common Japanese name for Japan (日本 nihon).

Very little is known about nihonium, as it has only been made in very small amounts that decay away within
seconds. The anomalously long lives of some superheavy nuclides, including some nihonium isotopes, are
explained by the "island of stability" theory. Experiments support the theory, with the half-lives of the
confirmed nihonium isotopes increasing from milliseconds to seconds as neutrons are added and the island
is approached. Nihonium has been calculated to have similar properties to its homologues boron,
aluminium, gallium, indium, and thallium. All but boron are post-transition metals, and nihonium is expected
to be a post-transition metal as well. It should also show several major differences from them; for example,
nihonium should be more stable in the +1 oxidation state than the +3 state, like thallium, but in the +1 state
nihonium should behave more like silver and astatine than thallium. Preliminary experiments in 2017
showed that elemental nihonium is not very volatile; its chemistry remains largely unexplored.

Contents
1History
1.1Early indications
1.2JINR–LLNL collaboration
1.3Riken
1.4Road to confirmation
1.4.12004–2008
1.4.22009–2015
1.5Approval of discoveries
1.6Naming
2Isotopes
2.1Stability and half-lives
3Predicted properties
3.1Physical and atomic
3.2Chemical
4Experimental chemistry
5See also
6Notes
7References
8External links

History[edit]

See also: Discoveries of the chemical elements

Early indications[edit]
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
[a]
fusion 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

[12][13]
2003, bombarding bismuth-209 with zinc-70, but were unsuccessful both times.

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 Lthe ivermore, California, United States, they made
an attempt on element 114 (which was predicted to be a magic number, closing a proton shell, and more

[12]
stable than element 113).

In 1998, the JINR–LLNL collaboration started their attempt on element 114, bombarding a target of

[12]
plutonium-244 with ions of calcium-48:

244
94Pu
+ 48
20Ca
292 290
→ 114* → 114 + 2
n
− 290
+e → 113 + νe

289
A single atom was observed which was thought to be the isotope 114: the results were published in

[14]
January 1999. Despite numerous attempts to repeat this reaction, an isotope with these decay

[15]
properties has never again been found, and the exact identity of this activity is unknown. A 2016 paper
considered that the most likely explanation of the 1998 result is that two neutrons were emitted by the

290 290
produced compound nucleus, leading to 114 and electron capture to 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

[7]
of element 113, but it was not recognised at the time, and the assignment is still uncertain. A similar long-

242 48
lived activity observed by the JINR team in March 1999 in the Pu + Ca reaction may be due to the

287 287 [8]


electron-capture daughter of 114, 113; this assignment is also tentative.
JINR–LLNL collaboration[edit]
The now-confirmed discovery of element 114 was made in June 1999 when the JINR team repeated the

244 48 [16][17]
first Pu + Ca reaction from 1998; following this, the JINR team used the same hot fusion

248 48
technique to synthesise elements 116 and 118 in 2000 and 2002 respectively via the Cm + Ca and

249 48
Cf + 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

[12][18]
chains.

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

[18]
projectiles. The JINR–LLNL collaboration published its results in February 2004:

243
95Am
+ 48
20Ca
291 288
→ 115* → 115 + 3
n
284
→ 113 +
α
243
95Am
+ 48
20Ca
291 287
→ 115* → 115 + 4
n
283
→ 113 +
α
Four further alpha decays were observed, ending with the spontaneous fission of isotopes of element 105,

[18]
dubnium.

Riken[edit]
Kōsuke Morita and Hiroshi Matsumoto, celebrating the naming on 1 December 2016.

48
While the JINR–LLNL collaboration had been studying fusion reactions with 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

209 70
same Bi + 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

[19] 278
much more certain, and would not require the use of radioactive targets. In particular, the isotope 113

266
expected to be produced in this reaction would decay to the known Bh, which had been synthesised in

[20]
2000 by a team at the Lawrence Berkeley National Laboratory (LBNL) in Berkeley.

209 70 [21]
The bombardment of Bi with Zn at Riken began in September 2003. The team detected a single

278 [22]
atom of 113 in July 2004 and published their results that September:

209
83Bi
+ 70
30Zn
279 278
→ 113* → 113 +
n
278 274
The Riken team observed four alpha decays from 113, creating a decay chain passing through Rg,

270 266 262 [22]


Mt, and Bh before terminating with the spontaneous fission of Db. The decay data they

266
observed for the alpha decay of Bh matched the 2000 data, lending support for their claim.

262
Spontaneous fission of its daughter Db had not been previously known; the American team had
[20]
observed only alpha decay from this nuclide.

Road to confirmation[edit]
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

[23]
out to confirm their results by these methods.

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.
2004–2008[edit]

In June 2004 and again in December 2005, the JINR–LLNL collaboration strengthened their claim for the

268 288
discovery of element 113 by conducting chemical experiments on Db, the final decay product of 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

268
case for their claim, since the chemistry of dubnium is known. 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

[1][24] 268
group 5). Both the half-life and decay mode were confirmed for the proposed Db which lends

[24][25]
support to the assignment of the parent and daughter nuclei to elements 115 and 113 respectively.

[20]
Further experiments at the JINR in 2005 confirmed the observed decay data.

205 70
In November and December 2004, the Riken team studied the Tl + Zn reaction, aiming the zinc beam

274
onto a thalliumrather than a bismuth target, in an effort to directly produce Rg in a cross-bombardment

278
as it is the immediate daughter of 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

[26] 209
thallium has a higher melting point than bismuth. The Riken team then repeated the original Bi +

70 278
Zn reaction and produced a second atom of 113 in April 2005, with a decay chain that again

262
terminated with the spontaneous fission of 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

[20]
states.

243 26
In 2006, a team at the Heavy Ion Research Facility in Lanzhou, China, investigated the Am + Mg

266 262
reaction, producing four atoms of Bh. All four chains started with an alpha decay to Db; three chains

278
ended there with spontaneous fission, as in the 113 chains observed at Riken, while the remaining one

258 266 [23]


continued via another alpha decay to Lr, as in the Bh chains observed at LBNL.

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
285 282
→ 113* → 113 + 3
n
282
Two atoms of 113 were detected. The aim of this experiment had been to synthesise the isotopes

281 282 283 284


113 and 113 that would fill in the gap between isotopes produced via hot fusion ( 113 and 113)

278
and cold fusion ( 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

266
first decay chain ended in fission after four alpha decays, presumably originating from Db or its electron-

266
capture daughter Rf. Spontaneous fission was not observed in the second chain even after four alpha

266
decays. A fifth alpha decay in each chain could have been missed, since Db can theoretically undergo

262 262
alpha decay, in which case the first decay chain would have ended at the known Lr or No and the

258
second might have continued to the known long-lived 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-

[27]
known main body of the chart of nuclides.

2009–2015[edit]

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 group 4 and group 5 elements by their chemical

[20]
properties with enough confidence to allow this assignment. 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

[20]
had not been fulfilled. 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

[20]
to known isotopes.

266 248 23
In early 2009, the Riken team synthesised the decay product Bh directly in the Cm + Na reaction

278
to establish its link with 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
258 [28][29]
alpha decay to Lr.

249 48
In late 2009, the JINR–LLNL collaboration studied the Bk + 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

[12]
in Tennessee, United States, 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

[12]
periodic table. Two isotopes of element 117 were synthesised, decaying to element 115 and then

[30]
element 113:

249
97Bk
+ 48
20Ca
297 294
→ 117* → 117 + 3
n
290 286
→ 115 + α → 113 + α

249
97Bk
+ 48
20Ca
297 293
→ 117* → 117 + 4
n
289 285
→ 115 + α → 113 + α

285 286 282 283


The new isotopes 113 and 113 produced did not overlap with the previously claimed 113, 113,

284
and 113, so this reaction could not be used as a cross-bombardment to confirm the 2003 or 2006

[23]
claims.

274 205 70
In March 2010, the Riken team again attempted to synthesise Rg directly through the Tl + Zn

[26]
reaction with upgraded equipment; they failed again and abandoned this cross-bombardment route.

After 450 more days of irradiation of bismuth with zinc projectiles, Riken produced and identified another

278 [31]
113 atom in August 2012. In this case, a series of six alpha decays was observed, leading to an
isotope of mendelevium:

278
113 → 274
111Rg
+
α
→ 270
109Mt
+
α
→ 266
107Bh
+
α
→ 262
105Db
+
α
→ 258
103Lr
+
α
→ 254
101Md
+
α
262
This decay chain differed from the previous observations at Riken mainly in the decay mode of Db,
which was previously observed to undergo spontaneous fission, but in this case instead alpha decayed; the

262 258
alpha decay of Db to Lr is well-known. The team calculated the probability of accidental coincidence

−28 [31] 254 254


to be 10 , or totally negligible. The resulting Md atom then underwent electron capture to Fm,

250
which underwent the seventh alpha decay in the chain to the long-lived Cf, which has a half-life of

[32]
around thirteen years.

249 48
The Bk + Ca experiment was repeated at the JINR in 2012 and 2013 with consistent results, and

[23]
again at the GSI in 2014. In August 2013, a team of researchers at Lund University in Lund, Sweden,

243 48
and at the GSI announced that they had repeated the 2003 Am + Ca experiment, confirming the

[21][33]
findings of the JINR–LLNL collaboration. The same year, the 2003 experiment had been repeated at

289
the JINR, now also creating the isotope 115 that could serve as a cross-bombardment for confirming

293 285
their discovery of the element 117 isotope 117, as well as its daughter 113 as part of its decay chain.

[23] [34]
Further confirmation was published by the team at the LBNL in 2015.

Approval of discoveries[edit]
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
[35]
involving the JINR. 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

[36]
Riken being awarded credit for element 113 had been leaked to Japanese newspapers. For the first

[35]
time in history, a team of Asian physicists would name a new element. 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

[37]
position for the official publication of the JWP reports.

The full JWP reports were published in 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

278
chain of 113 were inconsistent, their sum was now confirmed to be consistent, strongly suggesting that

278 262
the initial and final states in 113 and its daughter Db were the same for all three events. The decay of

262 258 254 278


Db to Lr and Md was previously known, firmly anchoring the decay chain of 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

[23][26][38]
their daughter, but did not give this work shared credit.

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

[12][37]
synthesising superheavy elements at the JINR.

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

278
of the alpha decay energies measured in the 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

180 180m
state of Ta has a half-life of hours, but an excited state Ta 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

[39][40]
113. Two members of the JINR team published a journal article rebutting these criticisms against the

[41]
congruence of their data on elements 113, 115, and 117 in June 2017.

Naming[edit]
Using Mendeleev's nomenclature for unnamed and undiscovered elements, nihonium should be known as
eka-thallium. In 1979, IUPAC published recommendations according to which the element was to be called

[42]
ununtrium (with the corresponding symbol of Uut), 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, who called it "element 113", with the symbol of E113, (113), or

[1]
even simply 113.

Before the JWP recognition of their priority, the Japanese team had unofficially suggested various names:

[43]
japonium, after their home country; nishinanium, after Japanese physicist Yoshio Nishina, the "founding

[44] [43]
father of modern physics research in Japan"; and rikenium, after the institute. 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. Japoniumwas 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

[45]
two Japanese pronunciations for the name of Japan. The discoverers also intended to reference

[46]
the support of their research by the Japanese people (Riken being almost entirely government-funded),
recover lost pride and trust in science among those who were affected by the Fukushima Daiichi nuclear

[47]
disaster, and honour Japanese chemist Masataka Ogawa's 1908 discovery of rhenium, which he named

[38]
"nipponium" with symbol Np after the other Japanese pronunciation of Japan's name. As Ogawa's claim
had not been accepted, the name "nipponium" could not be reused for a new element, and its symbol Np

[b]
had since been used for neptunium. In March 2016, Morita proposed the name "nihonium" to IUPAC, with

[38]
the symbol Nh.

Wikinews has related


news:IUPAC proposes
four new chemical
element names

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–

[36] [47]
IUPAC press release. Thus, IUPAC and IUPAP publicised the proposal of nihonium that June, and set
a five-month term to collect comments, after which the name would be formally established at a

[50][51] [52]
conference. The name was officially approved in November 2016. The naming ceremony for the
new element was held in Tokyo, Japan, in March 2017, with Naruhito, Crown Prince of Japan, in

[53]
attendance.

Isotopes[edit]

Main article: Isotopes of nihonium

Nihonium has no stable or naturally occurring isotopes. Several radioactive isotopes have been
synthesised in the laboratory, either by fusing two atoms or by observing the decay of heavier elements.

287
Eight different isotopes of nihonium have been reported with atomic masses 278, 282–287, and 290 ( Nh

290 [54]
and Nh are unconfirmed); they all decay through alpha decay to isotopes of roentgenium; there have

[55]
been indications that nihonium-284 can also decay by electron capture to copernicium-284.

Stability and half-lives[edit]

A chart of heavy nuclides with their known and predicted half-lives (known nuclides shown with borders).
Nihonium (row 113) is expected to be within the "island of stability" (white circle) and thus its nuclei are
slightly more stable than would otherwise be predicted; the known nihonium isotopes are too neutron-
poor to be within the island.
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 protons, so that the strong nuclear force cannot hold the
nucleus together against spontaneous fission for long. Calculations suggest that in the absence of other
stabilising factors, elements with more than 103 protons should not exist. Researchers in the 1960s
suggested that the closed nuclear shellsaround 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

[56][57]
nuclides become longer-lived as they approach the predicted location of the island.

All nihonium isotopes are unstable and radioactive; the heavier nihonium isotopes are more stable than the

286
lighter ones, as they are closer to the centre of the island. The most stable known nihonium isotope, Nh,

285 287
is also the heaviest; it has a half-life of 8 seconds. The isotope Nh, as well as the unconfirmed Nh

290 284 283


and Nh, have also been reported to have half-lives of over a second. The isotopes Nh and Nh
have half-lives of 1 and 0.1 seconds respectively. The remaining two isotopes have half-lives between 0.1

282 278
and 100 milliseconds: Nh has a half-life of 70 milliseconds, and 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

[57][58]
through 113), where each extra neutron so far multiplies the half-life by a factor of 5 to 20.

Predicted properties[edit]
Physical and atomic[edit]
[59]
Atomic energy levels of outermost s, p, and d electrons of thallium and nihonium

Nihonium is the first member of the 7p series of elements and the heaviest group 13element 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

[5]
velocities close to the speed of light. 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

[60]
the other four. 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 more

[5][c]
and less stabilised parts of the 7p subshell, respectively. For theoretical purposes, the valence electron

2 1 [1]
configuration may be represented to reflect the 7p subshell split as 7s 7p1/2 . The first ionisation energy

[1]
of nihonium is expected to be 7.306 eV, the highest among the metals of group 13. Similar subshell

splitting should exist for the 6d electron levels, with four being 6d 3/2 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
[60]
allow for the possibility of exotic nihonium compounds without lighter group 13 analogues.

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 7p 1/2
orbitals. Thus nihonium is expected to be much denser than thallium, with a predicted density of about 16

3 3
to 18 g/cm compared to thallium's 11.85 g/cm , since nihonium atoms are heavier than thallium atoms but

[1][59]
have the same volume. Bulk nihonium is expected to have a hexagonal close-packedcrystal structure,

[6]
like thallium. The melting and boiling points of nihonium have been predicted to be 430 °C and 1100 °C

[1][2]
respectively, exceeding the values for gallium, indium, and thallium, following periodic trends.

Chemical[edit]
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-

[61]
shell elements (copernicium and flerovium), an unprecedented situation in the periodic table. Nihonium
is expected to be less reactive than thallium, because of the greater stabilisation and resultant chemical

[4]
inactivity of the 7s subshell in nihonium compared to the 6s subshell in thallium. The standard electrode

+
potential for the Nh /Nh couple is predicted to be 0.6 V, so nihonium should be a rather noble metal, as

[4]
unreactive as rhodium and ruthenium.

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

[1]
have +1 as its most stable oxidation state.

The simplest possible nihonium compound is the monohydride, NhH. The bonding is provided by the 7p 1/2
electron of nihonium and the 1s electron of hydrogen. The SO interaction causes the binding energy of

[1]
nihonium monohydride to be reduced by about 1 eV and the nihonium–hydrogen bond length to

decrease as the bonding 7p1/2 orbital is relativistically contracted. This is unique among the 7p element

[62]
monohydrides; all the others have relativistic expansion of the bond length instead of contraction.
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

[63] [59]
thallium monohydride (TlH). The analogous monofluoride (NhF) should also exist. Nihonium(I) is

[1] +
predicted to be more similar to silver(I) than thallium(I): 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 Nh2O, which would be soluble in aqueous ammonia and

[4]
weakly soluble in water.

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

2 2 2 6
effectively leads to a valence shell closing at the 7s 7p configuration rather than the expected 7s 7p
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

2 2 [64]
three valence electrons outside the 7s 7p closed shell. ) 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

[5]
thallium as the SO splitting of the 7p subshell is greater than that for the 6p subshell. Nihonium should be

[1]
the most electronegative of the metallic group 13 elements, even more electronegative than tennessine,
the period 7 congener of the halogens: in the compound NhTs, the negative charge is expected to be on

[59]
the nihonium atom rather than the tennessine atom. The −1 oxidation should be more stable for

[1][65]
nihonium than for tennessine. The electron affinity of nihonium is calculated to be around 0.68 eV,

[1]
higher than thallium's at 0.4 eV; tennessine's is expected to be 1.8 eV, the lowest in its group. It is
theoretically predicted that nihonium should have an enthalpy of sublimation around 150 kJ/mol and an

[66]
enthalpy of adsorption on a gold surface around −159 kJ/mol.
BCl

3 has a trigonal structure.

NhCl

3 is predicted to be T-shaped.

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

[61]
metal character for nihonium. On the basis of the small energy gap between the 6d and 7s electrons,

[1][4]
the higher oxidation states +3 and +5 have been suggested for nihonium. Some simple compounds

with nihonium in the +3 oxidation state would be the trihydride (NhH 3), trifluoride (NhF3), and trichloride

(NhCl3). These molecules are predicted to be T-shaped and not trigonal planar as their boron analogues

[d] [63][e]
are: this is due to the influence of the 6d5/2 electrons on the bonding. 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

2
large 7s–7p energy gap means that they show reduced sp hybridisation compared to their boron

[63]
analogues.

The bonding in the lighter NhX3 molecules can be considered as that of a linear NhX+

2 species (similar to HgF2 or AuF−


2) 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 X 2 molecule and reduction

[63]
to nihonium(I):

NhX3 → NhX + X2

Nihonium thus continues the trend down group 13 of reduced stability of the +3 oxidation state, as all five of
[f]
these compounds have lower reaction energies than the unknown thallium(III) iodide. 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−
[63]
4.

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 X 2 molecule and reduction to
nihonium(III). Despite its instability, the possible existence of nihonium pentafluoride is entirely due to
relativistic effects allowing the 6d electrons to participate in the bonding. 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

[63]
trifluoride and pentafluoride.

Experimental chemistry[edit]
[66][71]
The chemical characteristics of nihonium have yet to be determined unambiguously. The isotopes

284 285 286 [66]


Nh, Nh, and Nh have half-lives long enough for chemical investigation. From 2010 to 2012,
some preliminary chemical experiments were performed at the JINR to determine the volatility of nihonium.

284 288 243 48


The isotope Nh was investigated, made as the daughter of Mc produced in the Am+ 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

284
to twenty atoms of 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

[66]
efficiently pass through the PTFE capillaries. 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

285 286
not kinetically favoured, so the longer-lived isotopes Nh and Nh were considered more desirable for

[66][72]
future experiments.

284 285 243 48


A 2017 experiment at the JINR, producing Nh and Nh via the Am+ Ca reaction as the daughters
288 289
of Mc and 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 a PTFE surface (−ΔHPTFE
ads(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
[73]
chromatography would be necessary to further probe the behaviour of elemental nihonium. 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

[74]
elements, including the superheavy dubnium.

S-ar putea să vă placă și