You are on page 1of 38

Wiki Loves Monuments: Photograph a monument, help Wikipedia and

win!Learn more

Flerovium
From Wikipedia, the free encyclopedia
Jump to navigationJump to search

Flerovium,  Fl 114

Flerovium

Pronunciation /flɪˈroʊviəm/[1] (flə-ROH-vee-əm)

Mass number [289] (unconfirmed: 290)

Flerovium in the periodic table

Hy
dro
ge
n
Lit Be
hiu ryll
m iu
m
So Ma
diu gn
m esi
um
Po CalSc
tas ciu an
siu m diu
m m

Ru Str Ytt
bid ont riu
iu iu m
m m

Ca Ba La C PrasNe ProSa Eu Ga Te Dy Ho Er ThuYt Lut


esi riu nth eri eod od me ma rop dol rbi spr lmi bi liu ter eti
um m an u ymi ym thi riu iu ini umosi um u m bi um
um m um iu um m m um um m u
m m

FraRa Ac T Prot Ur Ne Pl A Cu Be Cal Ein Fe Me NoLa


nci diu tiniho acti ani ptu ut me riu rkeiforstei rmnde belwr
um m um ri niu um niu on rici m liu niu niu iu levi iu enc
u m m iu um m m m m um m iu
m m m

nihonium ← flerovium

Atomic number (Z) 114

Group group 14 (carbon group)

Period period 7

Block p-block

Element category   Unknown chemical properties, but

probably a post-transition metal; possibly

a metalloid[2]

Electron configuration [Rn] 5f14 6d10 7s2 7p2 (predicted)[3]

Electrons per shell 2, 8, 18, 32, 32, 18, 4 (predicted)

Physical properties

Phase at STP gas (predicted)[3]

Boiling point ~ 210 K (~ −60 °C, ~ −80 °F) [4][5]

Density when liquid 14 g/cm3 (predicted)[6]

(at m.p.)

Heat of vaporization 38 kJ/mol (predicted)[6]

Atomic properties
Oxidation states (0), (+1), (+2), (+4), (+6) (predicted)[3][6][7]

 1st: 832.2 kJ/mol (predicted)[8]
Ionization energies
 2nd: 1600 kJ/mol (predicted)[6]

 3rd: 3370 kJ/mol (predicted)[6]

 (more)

Atomic radius empirical: 180 pm (predicted)[3][6]

Covalent radius 171–177 pm (extrapolated)[9]

Other properties

Natural occurrence synthetic

Crystal structure face-centred cubic (fcc)

(predicted)[10]

CAS Number 54085-16-4

History

Naming after Flerov Laboratory of Nuclear

Reactions (itself named after Georgy

Flyorov)[11]

Discovery Joint Institute for Nuclear Research (JINR)

and Lawrence Livermore National

Laboratory (LLNL) (1999)

Main isotopes of flerovium

Isotope Abun- Half-life  Decay Pro-


dance (t1/2) mode duct
284
Fl[12][13] syn 2.5 ms SF
285
Fl[14] syn 0.10 s α 281
Cn
40% α 282
Cn
286
Fl syn 0.12 s
60% SF
α 283
Cn
287
Fl[15] syn 0.48 s
EC? 287
Nh
288
Fl syn 0.66 s α 284
Cn
289
Fl syn 1.9 s α 285
Cn
EC 290
Nh
290
Fl[16][17] syn 19 s?
α 286
Cn

 Category: Flerovium

 view

 talk

 edit

 | references

Flerovium is a superheavy artificial chemical element with the symbol Fl and atomic


number 114. It is an extremely radioactive synthetic element. The element is named
after the Flerov Laboratory of Nuclear Reactions of the Joint Institute for Nuclear
Research in Dubna, Russia, where the element was discovered in 1998. The name of
the laboratory, in turn, honours the Russian physicist Georgy
Flyorov (Флёров in Cyrillic, hence the transliteration of "yo" to "e"). The name was
adopted by IUPAC on 30 May 2012.
In the periodic table of the elements, it is a transactinide element in the p-block. It is a
member of the 7th period and is the heaviest known member of the carbon group; it is
also the heaviest element whose chemistry has been investigated. Initial chemical
studies performed in 2007–2008 indicated that flerovium was unexpectedly volatile for a
group 14 element;[18] in preliminary results it even seemed to exhibit properties similar to
those of the noble gases.[19] More recent results show that flerovium's reaction
with gold is similar to that of copernicium, showing that it is a very volatile element that
may even be gaseous at standard temperature and pressure, that it would
show metallic properties, consistent with it being the heavier homologue of lead, and
that it would be the least reactive metal in group 14. The question of whether flerovium
behaves more like a metal or a noble gas is still unresolved as of 2018.
About 90 atoms of flerovium have been observed: 58 were synthesized directly, and the
rest were made from the radioactive decay of heavier elements. All of these flerovium
atoms have been shown to have mass numbers from 284 to 290. The most stable
known flerovium isotope, flerovium-289, has a half-life of around 1.9 seconds, but it is
possible that the unconfirmed flerovium-290 with one extra neutron may have a longer
half-life of 19 seconds; this would be one of the longest half-lives of any isotope of any
element at these farthest reaches of the periodic table. Flerovium is predicted to be near
the centre of the theorized island of stability, and it is expected that heavier flerovium
isotopes, especially the possibly doubly magic flerovium-298, may have even longer
half-lives.

Contents

 1Introduction
 2History
o 2.1Pre-discovery
o 2.2Discovery
o 2.3Isotopes
o 2.4Naming
 3Predicted properties
o 3.1Nuclear stability and isotopes
o 3.2Atomic and physical
o 3.3Chemical
 4Experimental chemistry
 5See also
 6Notes
 7References
 8Bibliography
 9Bibliography
 10External links

Introduction[edit]
This section is transcluded from Introduction to the heaviest elements.  (edit | history)
See also: Superheavy element §  Introduction

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

External video

 Visualization of unsuccessful nuclear

fusion, based on calculations by


the Australian National University[20]

The heaviest[a] atomic nuclei are created in nuclear reactions that combine two other
nuclei of unequal size[b] into one; roughly, the more unequal the two nuclei in terms of
mass, the greater the possibility that the two react. [26] The material made of the heavier
nuclei is made into a target, which is then bombarded by the beam of lighter nuclei. Two
nuclei can only fuse into one if they approach each other closely enough; normally,
nuclei (all positively charged) repel each other due to electrostatic repulsion. The strong
interaction can overcome this repulsion but only within a very short distance from a
nucleus; beam nuclei are thus greatly accelerated in order to make such repulsion
insignificant compared to the velocity of the beam nucleus. [27] Coming close alone is not
enough for two nuclei to fuse: when two nuclei approach each other, they usually
remain together for approximately 10−20 seconds and then part ways (not necessarily in
the same composition as before the reaction) rather than form a single nucleus. [27][28] If
fusion does occur, the temporary merger—termed a compound nucleus—is an excited
state. To lose its excitation energy and reach a more stable state, a compound nucleus
either fissions or ejects one or several neutrons,[c] which carry away the energy. This
occurs in approximately 10−16 seconds after the initial collision.[29][d]
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. [32] 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. [32] The transfer takes about
10−6 seconds; in order to be detected, the nucleus must survive this long. [35] The nucleus
is recorded again once its decay is registered, and the location, the energy, and the
time of the decay are measured.[32]
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, as it has unlimited range. [36] Nuclei of the heaviest elements
are thus theoretically predicted[37] and have so far been observed[38] to primarily decay via
decay modes that are caused by such repulsion: alpha decay and spontaneous fission;
[f]
 these modes are predominant for nuclei of superheavy elements. 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 determined arithmetically.
[g]
 Spontaneous fission, however, produces various nuclei as products, so the original
nuclide cannot be determined from its daughters. [h]
The information available to physicists aiming to synthesize one of the heaviest
elements 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. [i]

History[edit]
See also: Discovery of the chemical elements
Pre-discovery[edit]
From the late 1940s to the early 1960s, the early days of the synthesis of heavier and
heavier transuranium elements, it was predicted that since such heavy elements did not
occur naturally, they would have shorter and shorter half-lives to spontaneous fission,
until they stopped existing altogether at around element 108 (now known as hassium).
Initial work in the synthesis of the actinides appeared to confirm this.[50] The nuclear shell
model, introduced in 1949 and extensively developed in the late 1960s by William
Myers and Władysław Świątecki, stated that the protons and neutrons formed shells
within a nucleus, somewhat analogous to electrons forming electron shells within an
atom. The noble gases are unreactive due to their having full electron shells; thus it was
theorized that elements with full nuclear shells – having so-called "magic" numbers of
protons or neutrons – would be stabilized against radioactive decay. A doubly
magic isotope, having magic numbers of both protons and neutrons, would be
especially stabilized. Heiner Meldner calculated in 1965 that the next doubly magic
isotope after lead-208 would be flerovium-298 with 114 protons and 184 neutrons,
which would form the centre of a so-called "island of stability".[50][51] This island of stability,
supposedly ranging from copernicium (element 112) to oganesson (118), would come
after a long "sea of instability" from elements 101 (mendelevium) to 111 (roentgenium),
[50]
 and the flerovium isotopes in it were speculated in 1966 to have half-lives in excess of
a hundred million years.[52] These early predictions fascinated researchers, and led to the
first attempted synthesis of flerovium in 1968 using the reaction 248Cm(40Ar,xn). No
isotopes of flerovium were found in this reaction. This was thought to occur because the
compound nucleus 288Fl only has 174 neutrons instead of the hypothesized magic 184,
and this would have a significant impact on the reaction cross section (yield) and the
half-lives of nuclei produced.[53][54] It then took thirty more years for the first isotopes of
flerovium to be synthesized.[50] More recent work suggests that the local islands of
stability around hassium and flerovium are due to these nuclei being respectively
deformed and oblate, which make them resistant to spontaneous fission, and that the
true island of stability for spherical nuclei occurs at around unbibium-306 (with 122
protons and 184 neutrons).[55]
Discovery[edit]
Flerovium was first synthesized in December 1998 by a team of scientists at the Joint
Institute for Nuclear Research (JINR) in Dubna, Russia, led by Yuri Oganessian, who
bombarded a target of plutonium-244 with accelerated nuclei of calcium-48:
244

94Pu
 + 48
20Ca

 → 292
114Fl
* → 290
114Fl

 + 2 1
0n

This reaction had been attempted before, but without


success; for this 1998 attempt, the JINR had upgraded all
of its equipment to detect and separate the produced
atoms better and bombard the target more intensely. [56] A
single atom of flerovium, decaying by alpha emission with a
lifetime of 30.4 seconds, was detected. The decay
energy measured was 9.71 MeV, giving an expected half-
life of 2–23 s.[57] This observation was assigned to the
isotope flerovium-289 and was published in January 1999.
[57]
 The experiment was later repeated, but an isotope with
these decay properties was never found again and hence
the exact identity of this activity is unknown. It is possible
that it was due to the metastable isomer 289mFl,[58][59] but
because the presence of a whole series of longer-lived
isomers in its decay chain would be rather doubtful, the
most likely assignment of this chain is to the 2n channel
leading to 290Fl and electron capture to 290Nh, which fits well
with the systematics and trends across flerovium isotopes,
and is consistent with the low beam energy that was
chosen for that experiment, although further confirmation
would be desirable via the synthesis of 294Lv in
the 248Cm(48Ca,2n) reaction, which would alpha decay to 290Fl.
[16]
 The team at RIKEN reported a possible synthesis of the
isotopes 294Lv and 290Fl in 2016 through the 248Cm(48Ca,2n)
reaction, but the alpha decay of 294Lv was missed, alpha
decay of 290Fl to 286Cn was observed instead of electron
capture to 290Nh, and the assignment to 294Lv instead of 293Lv
and decay to an isomer of 285Cn was not certain.[17]
Glenn T. Seaborg, a scientist at the Lawrence Berkeley
National Laboratory who had been involved in work to
synthesize such superheavy elements, had said in
December 1997 that "one of his longest-lasting and most
cherished dreams was to see one of these magic
elements";[50] he was told of the synthesis of flerovium by his
colleague Albert Ghiorso soon after its publication in 1999.
Ghiorso later recalled:[60]
I wanted Glenn to know, so I went to his bedside and told
him. I thought I saw a gleam in his eye, but the next day
when I went to visit him he didn't remember seeing me. As
a scientist, he had died when he had that stroke. [60]

— Albert Ghiorso
Seaborg died two months later, on 25 February 1999. [60]
Isotopes[edit]
Main article: Isotopes of flerovium

List of flerovium isotopes

Half-life[j]
Decay Discovery Discovery
Isotope
mode year[61] reaction[62]
Value Ref

Pu(48Ca,4n)
240
Fl
284
2.5 ms [13]
SF 2015
Pu(48Ca,3n)
239

Fl
285
0.10 s [14]
α 2010 Pu(48Ca,5n)
242

Fl
286
0.12 s [63]
α, SF 2003 Lv(—,α)
290

Fl
287
0.48 s [63]
α, EC? 2003 Pu(48Ca,5n)
244

Fl
288
0.66 s [63]
α 2004 Pu(48Ca,4n)
244

Fl
289
1.9 s [63]
α 1999 Pu(48Ca,3n)
244

289m
Fl[k] 1.1 s [61]
α 2012 293m
Lv(—,α)

Fl[l]
290
19 s [16][17]
α, EC? 1998 Pu(48Ca,2n)
244

In March 1999, the same team replaced the 244Pu target


with a 242Pu one in order to produce other flerovium
isotopes. In this reaction, two atoms of flerovium were
produced, decaying via alpha emission with a half-life of
5.5 s. They were assigned as 287Fl.[64] This activity has not
been seen again either, and it is unclear what nucleus was
produced. It is possible that it was the meta-stable
isomer 287mFl[65] or the result of an electron capture branch
of 287Fl leading to 287Nh and 283Rg.[15]
The now-confirmed discovery of flerovium was made in
June 1999 when the Dubna team repeated the first reaction
from 1998. This time, two atoms of flerovium were
produced; they alpha decayed with a half-life of 2.6 s,
different from the 1998 result.[58] This activity was initially
assigned to 288Fl in error, due to the confusion regarding the
previous observations that were assumed to come
from 289Fl. Further work in December 2002 finally allowed a
positive reassignment of the June 1999 atoms to 289Fl.[65]
In May 2009, the Joint Working Party (JWP)
of IUPAC published a report on the discovery of
copernicium in which they acknowledged the discovery of
the isotope 283Cn.[66] This implied the discovery of flerovium,
from the acknowledgement of the data for the synthesis
of 287Fl and 291Lv, which decay to 283Cn. The discovery of the
isotopes flerovium-286 and -287 was confirmed in January
2009 at Berkeley. This was followed by confirmation of
flerovium-288 and -289 in July 2009 at the Gesellschaft für
Schwerionenforschung (GSI) in Germany. In 2011, IUPAC
evaluated the Dubna team experiments of 1999–2007.
They found the early data inconclusive, but accepted the
results of 2004–2007 as flerovium, and the element was
officially recognized as having been discovered. [67]
While the method of chemical characterisation of a
daughter was successful in the cases of flerovium and
livermorium, and the simpler structure of even–even
nuclei made the confirmation of oganesson (element 118)
straightforward, there have been difficulties in establishing
the congruence of decay chains from isotopes with odd
protons, odd neutrons, or both.[68][69] To get around this
problem with hot fusion, the decay chains from which
terminate in spontaneous fission instead of connecting to
known nuclei as cold fusion allows, experiments were
performed at Dubna in 2015 to produce lighter isotopes of
flerovium in the reactions of 48Ca with 239Pu and 240Pu,
particularly 283Fl, 284Fl, and 285Fl; the last had previously been
characterised in the 242Pu(48Ca,5n)285Fl reaction at
the Lawrence Berkeley National Laboratory in 2010. The
isotope 285Fl was more clearly characterised, while the new
isotope 284Fl was found to undergo immediate spontaneous
fission instead of alpha decay to known nuclides around
the N = 162 shell closure, and 283Fl was not found.[13] This
lightest isotope may yet conceivably be produced in the
cold fusion reaction 208Pb(76Ge,n)283Fl,[16] which the team
at RIKEN in Japan has considered investigating: [70][71] this
reaction is expected to have a higher cross-section of
200 fb than the "world record" low of 30 fb
for 209Bi(70Zn,n)278Nh, the reaction which RIKEN used for the
official discovery of element 113, now named nihonium.[16][72]
[73]
 The Dubna team repeated their investigation of
the 240Pu+48Ca reaction in 2017, observing three new
consistent decay chains of 285Fl, an additional decay chain
from this nuclide that may pass through some isomeric
states in its daughters, a chain that could be assigned
to 287Fl (likely stemming from 242Pu impurities in the target),
and some spontaneous fission events of which some could
be from 284Fl, though other interpretations including side
reactions involving the evaporation of charged particles are
also possible.[14]
Naming[edit]

Stamp of Russia, issued in 2013, dedicated to Georgy Flyorov and


flerovium

Using Mendeleev's nomenclature for unnamed and


undiscovered elements, flerovium is sometimes called eka-
lead. In 1979, IUPAC published recommendations
according to which the element was to be
called ununquadium (with the corresponding symbol
of Uuq),[74] a systematic element name as a placeholder,
until the discovery of the element is confirmed and a
permanent name is decided on. Most scientists in the field
called it "element 114", with the symbol
of E114, (114) or 114.[3]
According to IUPAC recommendations, the discoverer(s) of
a new element has the right to suggest a name. [75] After the
discovery of flerovium and livermorium was recognized by
IUPAC on 1 June 2011, IUPAC asked the discovery team
at the JINR to suggest permanent names for those two
elements. The Dubna team chose to name element
114 flerovium (symbol Fl),[76][77] after the Russian Flerov
Laboratory of Nuclear Reactions (FLNR), named after the
Soviet physicist Georgy Flyorov (also spelled Flerov);
earlier reports claim the element name was directly
proposed to honour Flyorov.[78] In accordance with the
proposal received from the discoverers IUPAC officially
named flerovium after the Flerov Laboratory of Nuclear
Reactions (an older name for the JINR), not after Flyorov
himself.[11] Flyorov is known for writing to Joseph Stalin in
April 1942 and pointing out the silence in scientific journals
in the field of nuclear fission in the United States, Great
Britain, and Germany. Flyorov deduced that this research
must have become classified information in those
countries. Flyorov's work and urgings led to the
development of the USSR's own atomic bomb project.
[77]
 Flyorov is also known for the discovery of spontaneous
fission with Konstantin Petrzhak. The naming ceremony for
flerovium and livermorium was held on 24 October 2012 in
Moscow.[79]
In a 2015 interview with Oganessian, the host, in
preparation to ask a question, said, "You said you had
dreamed to name [an element] after your teacher Georgy
Flyorov." Without letting the host finish, Oganessian
repeatedly said, "I did."[80]

Predicted properties[edit]
Very few properties of flerovium or its compounds have
been measured; this is due to its extremely limited and
expensive production[26] and the fact that it decays very
quickly. A few singular properties have been measured, but
for the most part, properties of flerovium remain unknown
and only predictions are available.
Nuclear stability and isotopes[edit]
Main article: Isotopes of flerovium
Regions of differently shaped nuclei, as predicted by the interacting
boson model[55]

The physical basis of the chemical periodicity governing the


periodic table is the electron shell closures at each noble
gas (atomic numbers 2, 10, 18, 36, 54, 86, and 118): as
any further electrons must enter a new shell with higher
energy, closed-shell electron configurations are markedly
more stable, leading to the relative inertness of the noble
gases.[6] Since protons and neutrons are also known to
arrange themselves in closed nuclear shells, the same
effect happens at nucleon shell closures, which happen at
specific nucleon numbers often dubbed "magic numbers".
The known magic numbers are 2, 8, 20, 28, 50, and 82 for
protons and neutrons, and additionally 126 for neutrons.
[6]
 Nucleons with magic proton and neutron numbers, such
as helium-4, oxygen-16, calcium-48, and lead-208, are
termed "doubly magic" and are very stable against decay.
This property of increased nuclear stability is very important
for superheavy elements: without any stabilization, their
half-lives would be expected by exponential extrapolation
to be in the range of nanoseconds (10−9 s) when element
110 (darmstadtium) is reached, because of the ever-
increasing repulsive electrostatic forces between the
positively charged protons that overcome the limited-
range strong nuclear force that holds the nucleus together.
The next closed nucleon shells and hence magic numbers
are thought to denote the centre of the long-sought island
of stability, where the half-lives to alpha decay and
spontaneous fission lengthen again.[6]

Orbitals with high azimuthal quantum number are raised in energy,


eliminating what would otherwise be a gap in orbital energy
corresponding to a closed proton shell at element 114. This raises the
next proton shell to the region around element 120.[55]

Initially, by analogy with the neutron magic number 126, the


next proton shell was also expected to occur at element
126, too far away from the synthesis capabilities of the mid-
20th century to achieve much theoretical attention. In 1966,
new values for the potential and spin-orbit interaction in this
region of the periodic table[81] contradicted this and predicted
that the next proton shell would occur instead at element
114,[6] and that nuclides in this region would be as stable
against spontaneous fission as many heavy nuclei such as
lead-208.[6] The expected closed neutron shells in this
region were at neutron number 184 or 196, thus
making 298Fl and 310Fl candidates for being doubly magic.
[6]
 1972 estimates predicted a half-life of about a year
for 298Fl, which was expected to be near a large island of
stability with the longest half-life at 294Ds (1010 years,
comparable to that of 232Th).[6] After the synthesis of the first
isotopes of elements 112 through 118 at the turn of the
21st century, it was found that the synthesized neutron-
deficient isotopes were stabilized against fission. In 2008 it
was thus hypothesized that the stabilization against fission
of these nuclides was due to their being oblate nuclei, and
that a region of oblate nuclei was centred on 288Fl.
Additionally, new theoretical models showed that the
expected gap in energy between the proton orbitals
2f7/2 (filled at element 114) and 2f5/2 (filled at element 120)
was smaller than expected, so that element 114 no longer
appeared to be a stable spherical closed nuclear shell. The
next doubly magic nucleus is now expected to be
around 306Ubb, but the expected low half-life and low
production cross section of this nuclide makes its synthesis
challenging.[55] Nevertheless, the island of stability is still
expected to exist in this region of the periodic table, and
nearer its centre (which has not been approached closely
enough yet) some nuclides, such as 291Mc and its alpha-
and beta-decay daughters,[m] may be found to decay
by positron emission or electron capture and thus move
into the centre of the island.[72] Due to the expected high
fission barriers, any nucleus within this island of stability
decays exclusively by alpha decay and perhaps some
electron capture and beta decay,[6] both of which would
bring the nuclei closer to the beta stability line where the
island is expected to be. Electron capture is needed to
reach the island, which is problematic because it is not
certain that electron capture becomes a major decay mode
in this region of the chart of nuclides.[72]
Several experiments have been performed between 2000
and 2004 at the Flerov Laboratory of Nuclear Reactions in
Dubna studying the fission characteristics of the compound
nucleus 292Fl by bombarding a plutonium-244 target with
accelerated calcium-48 ions.[82] A compound nucleus is a
loose combination of nucleons that have not yet arranged
themselves into nuclear shells. It has no internal structure
and is held together only by the collision forces between
the target and projectile nuclei.[83][n] The results revealed how
nuclei such as this fission predominantly by expelling
doubly magic or nearly doubly magic fragments such
as calcium-40, tin-132, lead-208, or bismuth-209. It was
also found that the yield for the fusion-fission pathway was
similar between calcium-48 and iron-58 projectiles,
indicating a possible future use of iron-58 projectiles in
superheavy element formation.[82] It has also been
suggested that a neutron-rich flerovium isotope can be
formed by the quasifission (partial fusion followed by
fission) of a massive nucleus.[84] Recently it has been shown
that the multi-nucleon transfer reactions in collisions of
actinide nuclei (such as uranium and curium) might be
used to synthesize the neutron-rich superheavy nuclei
located at the island of stability,[84] although production of
neutron-rich nobelium or seaborgium nuclei is more likely.[72]
Theoretical estimation of the alpha decay half-lives of the
isotopes of the flerovium supports the experimental data. [85]
[86]
 The fission-survived isotope 298Fl, long expected to be
doubly magic, is predicted to have alpha decay half-life
around 17 days.[87][88] The direct synthesis of the nucleus 298Fl
by a fusion–evaporation pathway is currently impossible
since no known combination of target and stable projectile
can provide 184 neutrons in the compound nucleus, and
radioactive projectiles such as calcium-50 (half-life
fourteen seconds) cannot yet be used in the needed
quantity and intensity.[84] Currently, one possibility for the
synthesis of the expected long-lived nuclei of copernicium
(291Cn and 293Cn) and flerovium near the middle of the island
include using even heavier targets such as curium-
250, berkelium-249, californium-251, and einsteinium-254,
that when fused with calcium-48 would produce nuclei such
as 291Mc and 291Fl (as decay products of 299Uue, 295Ts,
and 295Lv), with just enough neutrons to alpha decay to
nuclides close enough to the centre of the island to
possibly undergo electron capture and move inwards to the
centre, though the cross sections would be small and little
is yet known about the decay properties of superheavy
nuclides near the beta stability line. This may be the best
hope currently to synthesize nuclei on the island of stability,
but it is speculative and may or may not work in practice.
[72]
 Another possibility is to use controlled nuclear
explosions to achieve the high neutron flux necessary to
create macroscopic amounts of such isotopes.[72] This would
mimic the r-process in which the actinides were first
produced in nature and the gap of instability
after polonium bypassed, as it would bypass the gaps of
instability at 258–260Fm and at mass number 275 (atomic
numbers 104 to 108).[72] Some such isotopes
(especially 291Cn and 293Cn) may even have been
synthesized in nature, but would have decayed away far
too quickly (with half-lives of only thousands of years) and
be produced in far too small quantities (about 10 −12 the
abundance of lead) to be detectable as primordial
nuclides today outside cosmic rays.[72]
Atomic and physical[edit]
Flerovium is a member of group 14 in the periodic table,
below carbon, silicon, germanium, tin, and lead. Every
previous group 14 element has four electrons in its valence
shell, forming a valence electron configuration of ns2np2. In
flerovium's case, the trend will be continued and the
valence electron configuration is predicted to be 7s 27p2;
[3]
 flerovium will behave similarly to its lighter congeners in
many respects. Differences are likely to arise; a largely
contributing effect is the spin–orbit (SO) interaction—the
mutual interaction between the electrons' motion and spin.
It is especially strong for the superheavy elements,
because their electrons move faster than in lighter atoms,
at velocities comparable to the speed of light.[89] In relation
to flerovium atoms, it lowers the 7s and the 7p electron
energy levels (stabilizing the corresponding electrons), but
two of the 7p electron energy levels are stabilized more
than the other four.[90] The stabilization of the 7s electrons is
called the inert pair effect, and the effect "tearing" the 7p
subshell into the more stabilized and the less stabilized
parts is called subshell splitting. Computation 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 stabilized and less
stabilized parts of the 7p subshell, respectively. [91][o] For
many theoretical purposes, the valence electron
configuration may be represented to reflect the 7p subshell
split as 7s2

7p 2

1/2 .[3] These effects cause flerovium's chemistry to be


somewhat different from that of its lighter neighbours.
Due to the spin-orbit splitting of the 7p subshell being very
large in flerovium, and the fact that both flerovium's filled
orbitals in the seventh shell are stabilized relativistically, the
valence electron configuration of flerovium may be
considered to have a completely filled shell. Its
first ionization energy of 8.539 eV (823.9 kJ/mol) should be
the second-highest in group 14.[3] The 6d electron levels are
also destabilized, leading to some early speculations that
they may be chemically active, although newer work
suggests that this is unlikely.[6] Because this first ionisation
energy is higher than that of silicon and germanium, though
still lower than that for carbon, it has been suggested that
flerovium could be classified as a metalloid.[2]
The closed-shell electron configuration of flerovium results
in the metallic bonding in metallic flerovium being weaker
than in the preceding and following elements; thus,
flerovium is expected to have a low boiling point,[3] and has
recently been suggested to be possibly a gaseous metal,
similar to the predictions for copernicium, which also has a
closed-shell electron configuration.[55] The melting and
boiling points of flerovium were predicted in the 1970s to be
around 70 °C and 150 °C,[3] significantly lower than the
values for the lighter group 14 elements (those of lead are
327 °C and 1749 °C respectively), and continuing the trend
of decreasing boiling points down the group. Although
earlier studies predicted a boiling point of ~1000 °C or
2840 °C,[6] this is now considered unlikely because of the
expected weak metallic bonding in flerovium and that group
trends would expect flerovium to have a low sublimation
enthalpy.[3] Recent experimental indications have suggested
that the pseudo-closed shell configuration of flerovium
results in very weak metallic bonding and hence that
flerovium is probably a gas at room temperature with a
boiling point of around −60 °C.[4] Like mercury, radon,
and copernicium, but not lead and oganesson (eka-radon),
flerovium is calculated to have no electron affinity.[92]
In the solid state, flerovium is expected to be a dense metal
due to its high atomic weight, with a density variously
predicted to be either 22 g/cm3 or 14 g/cm3.[3] Flerovium is
expected to crystallize in the face-centred cubic crystal
structure like that of its lighter congener lead, [10] although
earlier calculations predicted a hexagonal close-
packed crystal structure due to spin-orbit coupling effects.
[93]
 The electron of the hydrogen-like flerovium ion (oxidized
so that it only has one electron, Fl113+) is expected to move
so fast that it has a mass 1.79 times that of a stationary
electron, due to relativistic effects. For comparison, the
figures for hydrogen-like lead and tin are expected to be
1.25 and 1.073 respectively.[94] Flerovium would form
weaker metal–metal bonds than lead and would
be adsorbed less on surfaces.[94]
Chemical[edit]
Flerovium is the heaviest known member of group 14 in the
periodic table, below lead, and is projected to be the
second member of the 7p series of chemical elements.
Nihonium and flerovium are expected to form a very short
subperiod, coming between the filling of the 6d 5/2 and
7p1/2 subshells. Their chemical behaviour is expected to be
very distinctive: nihonium's homology to thallium has been
called "doubtful" by computational chemists, while
flerovium's to lead has been called only "formal". [95]
The first five members of group 14 show the group
oxidation state of +4 and the latter members have an
increasingly prominent +2 chemistry due to the onset of the
inert pair effect. Tin represents the point at which the
stability of the +2 and +4 states are similar, and lead(II) is
the most stable of all the chemically well-understood group
14 elements in the +2 oxidation state. [3] The 7s orbitals are
very highly stabilized in flerovium and thus a very large
sp3 orbital hybridization is required to achieve the +4
oxidation state, so flerovium is expected to be even more
stable than lead in its strongly predominant +2 oxidation
state and its +4 oxidation state should be highly unstable.
[3]
 For example, flerovium dioxide (FlO2) is expected to be
highly unstable to decomposition into its constituent
elements (and would not be formed from the direct reaction
of flerovium with oxygen),[3][96] and flerovane (FlH4), which
should have Fl–H bond lengths of 1.787 Å,[7] is predicted to
be more thermodynamically unstable than plumbane,
spontaneously decomposing into flerovium(II) hydride
(FlH2) and hydrogen gas.[97] Flerovium tetrafluoride (FlF4)
[98]
 would have bonding mostly due to sd hybridizations
rather than sp3 hybridizations,[99] and its decomposition to
the difluoride and fluorine gas would be exothermic. [7] The
gross destabilization of all the tetrahalides (for example,
FlCl4 is destabilized by about 400 kJ/mol) is unfortunate
because otherwise these compounds would be very useful
in gas-phase chemical studies of flerovium.[7] The
corresponding polyfluoride anion FlF2−
6 should be unstable to hydrolysis in aqueous solution, and

flerovium(II) polyhalide anions such as FlBr−


3 and FlI

3 are predicted to form preferentially in flerovium-containing

solutions.[3] The sd hybridizations were suggested in early


calculations as the 7s and 6d electrons in flerovium share
approximately the same energy, which would allow a
volatile hexafluoride to form, but later calculations do not
confirm this possibility.[6] In general, the spin-orbit
contraction of the 7p1/2 orbital should lead to smaller bond
lengths and larger bond angles: this has been theoretically
confirmed in FlH2.[7] Nevertheless, even FlH2 should be
relativistically destabilized by 2.6 eV to below Fl+H2; the
large spin–orbit effects also break down the usual singlet–
triplet divide in the group 14 dihydrides. FlF 2 and FlCl2 are
predicted to be more stable than FlH2.[100]
Due to the relativistic stabilization of flerovium's 7s 27p
2

1/2  valence electron configuration, the 0 oxidation state


should also be more stable for flerovium than for lead, as
the 7p1/2 electrons begin to also exhibit a mild inert pair
effect:[3] this stabilization of the neutral state may bring
about some similarities between the behaviour of flerovium
and the noble gas radon.[19] Due to the expected relative
inertness of flerovium, its diatomic compounds FlH and FlF
should have lower energies of dissociation than the
corresponding lead compounds PbH and PbF.
[7]
 Flerovium(IV) should be even more electronegative than
lead(IV);[98] lead(IV) has electronegativity 2.33 on the
Pauling scale, though the lead(II) value is only 1.87.
Flerovium is expected to be a noble metal.[3]
Flerovium(II) should be more stable than lead(II), and
polyhalide ions and compounds of types FlX+, FlX2, FlX−
3, and FlX
2−

4 (X = Cl, Br, I) are expected to form readily. The fluorides

would undergo strong hydrolysis in aqueous solution. [3] All


the flerovium dihalides are expected to be stable, [3] with the
difluoride being water-soluble.[101] Spin-orbit effects would
destabilize flerovium dihydride (FlH2) by almost 2.6 eV
(250 kJ/mol).[96] In solution, flerovium would also form
the oxyanion flerovite (FlO2−
2) in aqueous solution, analogous to plumbite. Flerovium(II)

sulfate (FlSO4) and sulfide (FlS) should be very insoluble in


water, and flerovium(II) acetate (FlC2H3O2) and nitrate
(Fl(NO3)2) should be quite water-soluble.[6] The standard
electrode potential for the reduction of Fl2+ ions to metallic
flerovium is estimated to be around +0.9 V, confirming the
increased stability of flerovium in the neutral state. [3] In
general, due to the relativistic stabilization of the
7p1/2 spinor, Fl2+ is expected to have properties intermediate
between those of Hg2+ or Cd2+ and its lighter congener Pb2+.[3]

Experimental chemistry[edit]
Flerovium is currently the heaviest element to have had its
chemistry experimentally investigated, although the
chemical investigations have so far not led to a conclusive
result. Two experiments were performed in April–May 2007
in a joint FLNR-PSI collaboration aiming to study the
chemistry of copernicium. The first experiment involved the
reaction 242Pu(48Ca,3n)287Fl and the second the
reaction 244Pu(48Ca,4n)288Fl: these reactions produce short-
lived flerovium isotopes whose copernicium daughters
would then be studied.[102] The adsorption properties of the
resultant atoms on a gold surface were compared with
those of radon, as it was then expected that copernicium's
full-shell electron configuration would lead to noble-gas like
behaviour.[102] Noble gases interact with metal surfaces very
weakly, which is uncharacteristic of metals.[102]
The first experiment allowed detection of three atoms
of 283Cn but also seemingly detected 1 atom of 287Fl. This
result was a surprise given the transport time of the product
atoms is ~2 s, so the flerovium atoms produced should
have decayed to copernicium before adsorption. In the
second reaction, 2 atoms of 288Fl and possibly 1 atom of 289Fl
were detected. Two of the three atoms displayed
adsorption characteristics associated with a volatile, noble-
gas-like element, which has been suggested but is not
predicted by more recent calculations. These experiments
provided independent confirmation for the discovery of
copernicium, flerovium, and livermorium via comparison
with published decay data. Further experiments in 2008 to
confirm this important result detected a single atom of 289Fl,
and supported previous data showing flerovium having a
noble-gas-like interaction with gold.[102]
The experimental support for a noble-gas-like flerovium
soon weakened. In 2009 and 2010, the FLNR-PSI
collaboration synthesized further atoms of flerovium to
follow up their 2007 and 2008 studies. In particular, the first
three flerovium atoms synthesized in the 2010 study
suggested again a noble-gas-like character, but the
complete set taken together resulted in a more ambiguous
interpretation, unusual for a metal in the carbon group but
not fully like a noble gas in character.[103] In their paper, the
scientists refrained from calling flerovium's chemical
properties "close to those of noble gases", as had
previously been done in the 2008 study.[103] Flerovium's
volatility was again measured through interactions with a
gold surface, and provided indications that the volatility of
flerovium was comparable to that of mercury, astatine, and
the simultaneously investigated copernicium, which had
been shown in the study to be a very volatile noble metal,
conforming to its being the heaviest group 12 element
known.[103] Nevertheless, it was pointed out that this volatile
behaviour was not expected for a usual group 14 metal. [103]
In even later experiments from 2012 at the GSI, the
chemical properties of flerovium were found to be more
metallic than noble-gas-like. Jens Volker Kratz and
Christoph Düllmann specifically named copernicium and
flerovium as belonging to a new category of "volatile
metals"; Kratz even speculated that they might be gaseous
at standard temperature and pressure.[55][104] These "volatile
metals", as a category, were expected to fall between
normal metals and noble gases in terms of adsorption
properties.[55] Contrary to the 2009 and 2010 results, it was
shown in the 2012 experiments that the interactions of
flerovium and copernicium respectively with gold were
about equal.[105] Further studies showed that flerovium was
more reactive than copernicium, in contradiction to
previous experiments and predictions. [55]
In a 2014 paper detailing the experimental results of the
chemical characterisation of flerovium, the GSI group
wrote: "[flerovium] is the least reactive element in the
group, but still a metal."[106] Nevertheless, in a 2016
conference about the chemistry and physics of heavy and
superheavy elements, Alexander Yakushev and Robert
Eichler, two scientists who had been active at GSI and
FLNR in determining the chemistry of flerovium, still urged
caution based on the inconsistencies of the various
experiments previously listed, noting that the question of
whether flerovium was a metal or a noble gas was still
open with the available evidence: one study suggested a
weak noble-gas-like interaction between flerovium and
gold, while the other suggested a stronger metallic
interaction. The same year, new experiments aimed at
probing the chemistry of copernicium and flerovium were
conducted at GSI's TASCA facility, and the data from these
experiments is currently being analysed. As such,
unambiguous determination of the chemical characteristics
of flerovium has yet to have been established,[107] although
the experiments to date have allowed the first experimental
estimation of flerovium's boiling point: around −60 °C, so
that it is probably a gas at standard conditions. [4] The
longer-lived flerovium isotope 289Fl has been considered of
interest for future radiochemical studies.[108]

See also[edit]
 Island of stability: Flerovium–Unbinilium–
Unbihexium
 Isotopes of flerovium
 Extended periodic table
Portals
Access related topics  Chemistry portal

Find out more on  Media


Wikipedia's from Commons
Sister projects
 Definitions
from Wiktionary

Notes[edit]
1. ^ In nuclear physics, an element is called heavy if its
atomic number is high; lead(element 82) is one example
of such a heavy element. The term "superheavy
elements" typically refers to elements with atomic number
greater than 103 (although there are other definitions,
such as atomic number greater than 100[21] or 112;
[22]
 sometimes, the term is presented an equivalent to the
term "transactinide", which puts an upper limit before the
beginning of the hypothetical superactinide series).
[23]
 Terms "heavy isotopes" (of a given element) and
"heavy nuclei" mean what could be understood in the
common language—isotopes of high mass (for the given
element) and nuclei of high mass, respectively.
2. ^ In 2009, a team at 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.[24] In comparison, the reaction that
resulted in hassium discovery, 208Pb + 58Fe, had a cross
section of ~20 pb (more specifically, 19+19
−11 pb), as estimated by the discoverers.
[25]

3. ^ The greater the excitation energy, the more neutrons


are ejected. If the excitation energy is lower than energy
binding each neutron to the rest of the nucleus, neutrons
are not emitted; instead, the compound nucleus de-
excites by emitting a gamma ray.[29]
4. ^ The definition by the IUPAC/IUPAP Joint Working
Party states that a chemical elementcan only be
recognized as discovered if a nucleus of it has
not decayed within 10−14seconds. 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.[30] This figure also marks the generally
accepted upper limit for lifetime of a compound nucleus.[31]
5. ^ This separation is based on that the resulting nuclei
move past the target more slowly then the unreacted
beam nuclei. The separator contains electric and
magnetic fields whose effects on a moving particle cancel
out for a specific velocity of a particle.[33] 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.[34]
6. ^ Not all decay modes are caused by electrostatic
repulsion. For example, beta decay is caused by
the weak interaction.[39]
7. ^ 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 heaviest nuclei.[40] The first direct
measurement of mass of a superheavy nucleus was
reported in 2018 at LBNL.[41]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).[42]
8. ^ Spontaneous fission was discovered by Soviet
physicist Georgy Flerov,[43] a leading scientist at JINR,
and thus it was a "hobbyhorse" for the facility.[44] 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. [31] They thus
preferred to link new isotopes to the already known ones
by successive alpha decays.[43]
9. ^ For instance, element 102 was mistakenly identified in
1957 at the Nobel Institute of Physics
in Stockholm, Stockholm County, Sweden.[45] 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. [46] 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.[46]JINR
insisted that they were the first to create the element and
suggested a name of their own for the new
element, joliotium;[47] the Soviet name was also not
accepted (JINR later referred to the naming of
element 102 as "hasty").[48] The name "nobelium"
remained unchanged on account of its widespread
usage.[49]
10. ^ Different sources give different values for half-lives; the
most recently published values are listed.
11. ^ This isotope is unconfirmed
12. ^ This isotope is unconfirmed
13. ^ Specifically,  Mc,  Fl,  Nh,  Nh,  Cn,  Rg,  Rg,
291 291 291 287 287 287 283

and  Ds, which are expected to decay to the relatively


283

longer-lived nuclei 283Mt, 287Ds, and 291Cn.[72]


14. ^ It is estimated that it requires around 10  s for the−14

nucleons to arrange themselves into nuclear shells, at


which point the compound nucleus becomes a nuclide,
and this number is used by IUPAC as the minimum half-
life a claimed isotope must have to be recognized as a
nuclide.[83]
15. ^ The quantum number corresponds to the letter in the
electron orbital name: 0 to s, 1 to p, 2 to d, etc.
See azimuthal quantum number for more information.

References[edit]
1. ^ "Flerovium and Livermorium".  Periodic Table of
Videos.  The University of Nottingham. Retrieved  4
June  2012.
2. ^ Jump up to:a b Gong, Sheng; Wu, Wei; Wang, Fancy Qian;
Liu, Jie; Zhao, Yu; Shen, Yiheng; Wang, Shuo; Sun,
Qiang; Wang, Qian (8 February 2019). "Classifying
superheavy elements by machine learning".  Physical
Review A. 99: 022110-1–
7. doi:10.1103/PhysRevA.99.022110.
3. ^ Jump up to:a b c d e f g h i j k l m n o p q r s t u v Hoffman, Darleane
C.; Lee, Diana M.; Pershina, Valeria (2006).
"Transactinides and the future elements". In Morss;
Edelstein, Norman M.; Fuger, Jean (eds.).  The
Chemistry of the Actinide and Transactinide
Elements  (3rd ed.). Dordrecht, The
Netherlands: Springer Science+Business
Media. ISBN 978-1-4020-3555-5.
4. ^ Jump up to:a b c Oganessian, Yu. Ts. (27 January
2017). "Discovering Superheavy Elements".  Oak Ridge
National Laboratory. Retrieved 21 April  2017.
5. ^ Seaborg, G. T.  "Transuranium
element". Encyclopædia Britannica. Retrieved  16
March2010.
6. ^ Jump up to:a b c d e f g h i j k l m n o p q r Fricke, Burkhard
(1975).  "Superheavy elements: a prediction of their
chemical and physical properties". Recent Impact of
Physics on Inorganic Chemistry. Structure and
Bonding.  21: 89–
144.  doi:10.1007/BFb0116498. ISBN 978-3-540-07109-
9. Retrieved 4 October 2013.
7. ^ Jump up to:a b c d e f Schwerdtfeger, Peter; Seth, Michael
(2002).  "Relativistic Quantum Chemistry of the
Superheavy Elements. Closed-Shell Element 114 as a
Case Study"  (PDF).  Journal of Nuclear and
Radiochemical Sciences. 3  (1): 133–
136.  doi:10.14494/jnrs2000.3.133. Retrieved  12
September  2014.
8. ^ Pershina, Valeria. "Theoretical Chemistry of the
Heaviest Elements". In Schädel, Matthias; Shaughnessy,
Dawn (eds.). The Chemistry of Superheavy
Elements  (2nd ed.). Springer Science & Business Media.
p. 154.  ISBN  9783642374661.
9. ^ Bonchev, Danail; Kamenska, Verginia
(1981).  "Predicting the Properties of the 113–120
Transactinide Elements".  Journal of Physical Chemistry.
American Chemical Society.  85(9): 1177–
1186.  doi:10.1021/j150609a021.
10. ^ Jump up to:a b Maiz Hadj Ahmed, H.; Zaoui, A.; Ferhat, M.
(2017).  "Revisiting the ground state phase stability of
super-heavy element Flerovium".  Cogent
Physics. 4  (1).  doi:10.1080/23311940.2017.1380454.
Retrieved 26 November 2018.
11. ^ Jump up to:a b "Element 114 is Named Flerovium and
Element 116 is Named Livermorium"  (Press
release). IUPAC. 30 May 2012.
12. ^ Utyonkov, V.K. et al. (2015) Synthesis of superheavy
nuclei at limits of stability: 239,240Pu + 48Ca and 249–251Cf + 48Ca
reactions. Super Heavy Nuclei International Symposium,
Texas A & M University, College Station TX, USA, March
31 – April 02, 2015
13. ^ Jump up to:a b c Utyonkov, V. K.; Brewer, N. T.; Oganessian,
Yu. Ts.; Rykaczewski, K. P.; Abdullin, F. Sh.; Dmitriev, S.
N.; Grzywacz, R. K.; Itkis, M. G.; Miernik, K.; Polyakov,
A. N.; Roberto, J. B.; Sagaidak, R. N.; Shirokovsky, I. V.;
Shumeiko, M. V.; Tsyganov, Yu. S.; Voinov, A. A.;
Subbotin, V. G.; Sukhov, A. M.; Sabel'nikov, A. V.;
Vostokin, G. K.; Hamilton, J. H.; Stoyer, M. A.; Strauss,
S. Y. (15 September 2015). "Experiments on the
synthesis of superheavy nuclei  284Fl and  285Fl in
the  239,240Pu +  48Ca reactions".  Physical Review C. 92 (3):
034609-1–034609-10.  Bibcode:2015PhRvC..92c4609U. 
doi:10.1103/PhysRevC.92.034609.
14. ^ Jump up to:a b c Utyonkov, V. K.; Brewer, N. T.; Oganessian,
Yu. Ts.; Rykaczewski, K. P.; Abdullin, F. Sh.; Dimitriev,
S. N.; Grzywacz, R. K.; Itkis, M. G.; Miernik, K.;
Polyakov, A. N.; Roberto, J. B.; Sagaidak, R. N.;
Shirokovsky, I. V.; Shumeiko, M. V.; Tsyganov, Yu. S.;
Voinov, A. A.; Subbotin, V. G.; Sukhov, A. M.; Karpov, A.
V.; Popeko, A. G.; Sabel'nikov, A. V.; Svirikhin, A. I.;
Vostokin, G. K.; Hamilton, J. H.; Kovrinzhykh, N. D.;
Schlattauer, L.; Stoyer, M. A.; Gan, Z.; Huang, W. X.; Ma,
L. (30 January 2018). "Neutron-deficient superheavy
nuclei obtained in the  240Pu+48Ca reaction".  Physical
Review C.  97  (1): 014320–1—014320–
10.  Bibcode:2018PhRvC..97a4320U. doi:10.1103/PhysR
evC.97.014320.
15. ^ Jump up to:a b Hofmann, S.; Heinz, S.; Mann, R.; Maurer, J.;
Münzenberg, G.; Antalic, S.; Barth, W.; Burkhard, H. G.;
Dahl, L.; Eberhardt, K.; Grzywacz, R.; Hamilton, J. H.;
Henderson, R. A.; Kenneally, J. M.; Kindler, B.;
Kojouharov, I.; Lang, R.; Lommel, B.; Miernik, K.; Miller,
D.; Moody, K. J.; Morita, K.; Nishio, K.; Popeko, A. G.;
Roberto, J. B.; Runke, J.; Rykaczewski, K. P.; Saro, S.;
Schneidenberger, C.; Schött, H. J.; Shaughnessy, D. A.;
Stoyer, M. A.; Thörle-Pospiech, P.; Tinschert, K.;
Trautmann, N.; Uusitalo, J.; Yeremin, A. V. (2016).
"Remarks on the Fission Barriers of SHN and Search for
Element 120". In Peninozhkevich, Yu. E.; Sobolev, Yu.
G. (eds.).  Exotic Nuclei: EXON-2016 Proceedings of the
International Symposium on Exotic Nuclei. Exotic Nuclei.
pp.  155–164.  ISBN  9789813226555.
16. ^ Jump up to:a b c d e Hofmann, S.; Heinz, S.; Mann, R.;
Maurer, J.; Münzenberg, G.; Antalic, S.; Barth, W.;
Burkhard, H. G.; Dahl, L.; Eberhardt, K.; Grzywacz, R.;
Hamilton, J. H.; Henderson, R. A.; Kenneally, J. M.;
Kindler, B.; Kojouharov, I.; Lang, R.; Lommel, B.; Miernik,
K.; Miller, D.; Moody, K. J.; Morita, K.; Nishio, K.;
Popeko, A. G.; Roberto, J. B.; Runke, J.; Rykaczewski,
K. P.; Saro, S.; Scheidenberger, C.; Schött, H. J.;
Shaughnessy, D. A.; Stoyer, M. A.; Thörle-Popiesch, P.;
Tinschert, K.; Trautmann, N.; Uusitalo, J.; Yeremin, A. V.
(2016). "Review of even element super-heavy nuclei and
search for element 120".  The European Physics Journal
A.  2016  (52):
180.  Bibcode:2016EPJA...52..180H. doi:10.1140/epja/i2
016-16180-4.
17. ^ Jump up to:a b c Kaji, Daiya; Morita, Kosuke; Morimoto,
Kouji; Haba, Hiromitsu; Asai, Masato; Fujita, Kunihiro;
Gan, Zaiguo; Geissel, Hans; Hasebe, Hiroo; Hofmann,
Sigurd; Huang, MingHui; Komori, Yukiko; Ma, Long;
Maurer, Joachim; Murakami, Masashi; Takeyama, Mirei;
Tokanai, Fuyuki; Tanaka, Taiki; Wakabayashi, Yasuo;
Yamaguchi, Takayuki; Yamaki, Sayaka; Yoshida, Atsushi
(2017). "Study of the Reaction  48Ca +  248Cm →  296Lv* at
RIKEN-GARIS". Journal of the Physical Society of
Japan. 86 (3): 034201–1–
7. Bibcode:2017JPSJ...86c4201K. doi:10.7566/JPSJ.86.
034201.
18. ^ Eichler, Robert; et al. (2010). "Indication for a volatile
element 114"  (PDF). Radiochimica Acta.  98  (3): 133–
139.  doi:10.1524/ract.2010.1705. S2CID  95172228.
19. ^ Jump up to:a b Gäggeler, H. W. (5–7 November 2007). "Gas
Phase Chemistry of Superheavy Elements"  (PDF). Paul
Scherrer Institute. Archived from the original  (PDF) on 20
February 2012. Retrieved 10 August  2013.
20. ^ Wakhle, A.; Simenel, C.; Hinde, D. J.; et al. (2015).
Simenel, C.; Gomes, P. R. S.; Hinde, D. J.; et al.
(eds.). "Comparing Experimental and Theoretical
Quasifission Mass Angle Distributions". European
Physical Journal Web of Conferences. 86:
00061. Bibcode:2015EPJWC..8600061W. doi:10.1051/e
pjconf/20158600061. ISSN 2100-014X.
21. ^ Krämer, K. (2016). "Explainer: superheavy
elements".  Chemistry World. Retrieved  15 March 2020.
22. ^ "Discovery of Elements 113 and 115".  Lawrence
Livermore National Laboratory. Archived from  the
original on 11 September 2015. Retrieved  15
March  2020.
23. ^ Eliav, E.; Kaldor, U.; Borschevsky, A. (2018).
"Electronic Structure of the Transactinide Atoms". In
Scott, R. A. (ed.).  Encyclopedia of Inorganic and
Bioinorganic Chemistry.  John Wiley & Sons. pp.  1–
16.  doi:10.1002/9781119951438.eibc2632.  ISBN  978-1-
119-95143-8.
24. ^ Oganessian, Yu. Ts.; Dmitriev, S. N.; Yeremin, A. V.; et
al. (2009). "Attempt to produce the isotopes of element
108 in the fusion reaction  136Xe +  136Xe".  Physical Review
C.  79  (2):
024608. doi:10.1103/PhysRevC.79.024608.  ISSN  0556-
2813.
25. ^ Münzenberg, G.;  Armbruster, P.; Folger, H.; et al.
(1984).  "The identification of element
108"  (PDF). Zeitschrift für Physik A. 317 (2): 235–
236.  Bibcode:1984ZPhyA.317..235M.  doi:10.1007/BF01
421260. Archived from  the original  (PDF)  on 7 June
2015. Retrieved  20 October 2012.
26. ^ Jump up to:a b Subramanian, S. (2019).  "Making New
Elements Doesn't Pay. Just Ask This Berkeley
Scientist". Bloomberg Businessweek. Retrieved 18
January  2020.
27. ^ Jump up to:a b Ivanov, D. (2019). "Сверхтяжелые шаги в
неизвестное"  [Superheavy steps into the
unknown].  N+1 (in Russian). Retrieved 2
February  2020.
28. ^ Hinde, D. (2014). "Something new and superheavy at
the periodic table".  The Conversation. Retrieved  30
January  2020.
29. ^ Jump up to:a b Krása, A. (2010). "Neutron Sources for
ADS"  (PDF).  Czech Technical University in Prague.
pp.  4–8. Retrieved 20 October  2019.
30. ^ Wapstra, A. H.  (1991). "Criteria that must be satisfied
for the discovery of a new chemical element to be
recognized"  (PDF). Pure and Applied Chemistry. 63 (6):
883.  doi:10.1351/pac199163060879. ISSN 1365-3075.
Retrieved 28 August  2020.
31. ^ Jump up to:a b Hyde, E. K.;  Hoffman, D. C.; Keller, O. L.
(1987).  "A History and Analysis of the Discovery of
Elements 104 and 105".  Radiochimica Acta. 42 (2): 67–
68.  doi:10.1524/ract.1987.42.2.57.  ISSN  2193-3405.
32. ^ Jump up to:a b c Chemistry World (2016).  "How to Make
Superheavy Elements and Finish the Periodic Table
[Video]".  Scientific American. Retrieved  27
January  2020.
33. ^ Hoffman 2000, p. 334.
34. ^ Hoffman 2000, p. 335.
35. ^ Zagrebaev 2013, p. 3.
36. ^ Beiser 2003, p. 432.
37. ^ Staszczak, A.; Baran, A.; Nazarewicz, W. (2013).
"Spontaneous fission modes and lifetimes of superheavy
elements in the nuclear density functional
theory". Physical Review C.  87  (2): 024320–
1. arXiv:1208.1215. Bibcode:2013PhRvC..87b4320S. do
i:10.1103/physrevc.87.024320. ISSN 0556-2813.
38. ^ Audi 2017, pp. 030001-128–030001-138.
39. ^ Beiser 2003, p. 439.
40. ^ Oganessian, Yu. Ts.; Rykaczewski, K. P. (2015).  "A
beachhead on the island of stability". Physics
Today. 68 (8): 32–
38.  Bibcode:2015PhT....68h..32O.  doi:10.1063/PT.3.288
0. ISSN 0031-9228.  OSTI  1337838.
41. ^ Grant, A. (2018). "Weighing the heaviest
elements". Physics
Today. doi:10.1063/PT.6.1.20181113a.
42. ^ Howes, L. (2019).  "Exploring the superheavy elements
at the end of the periodic table". Chemical & Engineering
News. Retrieved  27 January 2020.
43. ^ Jump up to:a b Robinson, A. E. (2019). "The Transfermium
Wars: Scientific Brawling and Name-Calling during the
Cold War".  Distillations. Retrieved 22 February 2020.
44. ^ "Популярная библиотека химических элементов.
Сиборгий (экавольфрам)"[Popular library of chemical
elements. Seaborgium (eka-tungsten)]. n-t.ru  (in
Russian). Retrieved 7 January 2020. Reprinted
from "Экавольфрам" [Eka-tungsten]. Популярная
библиотека химических элементов. Серебро —
Нильсборий и далее [Popular library of chemical
elements. Silver through nielsbohrium and beyond] (in
Russian). Nauka. 1977.
45. ^ "Nobelium – Element information, properties and uses |
Periodic Table". Royal Society of Chemistry. Retrieved  1
March  2020.
46. ^ Jump up to:a b Kragh 2018, pp. 38–39.
47. ^ Kragh 2018, p. 40.
48. ^ Ghiorso, A.; Seaborg, G. T.; Oganessian, Yu. Ts.; et al.
(1993).  "Responses on the report 'Discovery of the
Transfermium elements' followed by reply to the
responses by Transfermium Working Group"  (PDF). Pure
and Applied Chemistry.  65  (8): 1815–
1824.  doi:10.1351/pac199365081815. Archived  (PDF)  fr
om the original on 25 November 2013. Retrieved  7
September  2016.
49. ^ Commission on Nomenclature of Inorganic Chemistry
(1997).  "Names and symbols of transfermium elements
(IUPAC Recommendations 1997)"  (PDF). Pure and
Applied Chemistry. 69 (12): 2471–
2474.  doi:10.1351/pac199769122471.
50. ^ Jump up to:a b c d e Sacks, O. (8 February 2004). "Greetings
From the Island of Stability".  The New York Times.
51. ^ Bemis, C.E.; Nix, J.R. (1977).  "Superheavy elements -
the quest in perspective"  (PDF). Comments on Nuclear
and Particle Physics.  7 (3): 65–78. ISSN 0010-2709.
52. ^ Emsley, John (2011).  Nature's Building Blocks: An A-Z
Guide to the Elements  (New ed.). New York, NY: Oxford
University Press. p. 580.  ISBN  978-0-19-960563-7.
53. ^ Hoffman, D.C; Ghiorso, A.; Seaborg, G.T. (2000). The
Transuranium People: The Inside Story. Imperial College
Press. Bibcode:2000tpis.book.....H.  ISBN  978-1-86094-
087-3.
54. ^ Epherre, M.; Stephan, C. (1975). "Les éléments
superlourds"  (PDF).  Le Journal de Physique
Colloques (in French).  11  (36): C5–159–
164.  doi:10.1051/jphyscol:1975541.
55. ^ Jump up to:a b c d e f g h Kratz, J. V. (5 September 2011). The
Impact of Superheavy Elements on the Chemical and
Physical Sciences  (PDF). 4th International Conference
on the Chemistry and Physics of the Transactinide
Elements. Retrieved 27 August  2013.
56. ^ Chapman, Kit (30 November 2016).  "What it takes to
make a new element".  Chemistry World. Royal Society of
Chemistry. Retrieved  3 December  2016.
57. ^ Jump up to:a b Oganessian, Yu. Ts.; et al.
(1999).  "Synthesis of Superheavy Nuclei in the  48Ca
+244Pu Reaction"  (PDF). Physical Review
Letters.  83  (16):
3154.  Bibcode:1999PhRvL..83.3154O. doi:10.1103/Phys
RevLett.83.3154.
58. ^ Jump up to:a b Oganessian, Yu. Ts.; et al.
(2000).  "Synthesis of superheavy nuclei in the  48Ca
+  244Pu reaction:  288114"  (PDF).  Physical Review
C.  62  (4):
041604. Bibcode:2000PhRvC..62d1604O. doi:10.1103/P
hysRevC.62.041604.
59. ^ 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"  (PDF). Physical
Review C.  70  (6):
064609. Bibcode:2004PhRvC..70f4609O.  doi:10.1103/P
hysRevC.70.064609. Archived from  the
original  (PDF)  on 28 May 2008.
60. ^ Jump up to:a b c Browne, M. W. (27 February 1999). "Glenn
Seaborg, Leader of Team That Found Plutonium, Dies at
86".  The New York Times. Archived from the original  on
22 May 2013. Retrieved  26 August 2013.
61. ^ Jump up to:a b Audi, G.; Kondev, F. G.; Wang, M.; Huang,
W. J.; Naimi, S. (2017).  "The NUBASE2016 evaluation of
nuclear properties"  (PDF). Chinese Physics C. 41 (3):
030001. Bibcode:2017ChPhC..41c0001A. doi:10.1088/1
674-1137/41/3/030001.
62. ^ Thoennessen, M. (2016). The Discovery of Isotopes: A
Complete Compilation. Springer. pp. 229, 234,
238.  doi:10.1007/978-3-319-31763-2.  ISBN  978-3-319-
31761-8.  LCCN  2016935977.
63. ^ Jump up to:a b c d Oganessian, Y.T. (2015).  "Super-heavy
element research". Reports on Progress in
Physics. 78 (3):
036301. Bibcode:2015RPPh...78c6301O. doi:10.1088/0
034-4885/78/3/036301.  PMID  25746203.
64. ^ Oganessian, Yu. Ts.; et al. (1999). "Synthesis of nuclei
of the superheavy element 114 in reactions induced
by  48Ca".  Nature.  400  (6741):
242.  Bibcode:1999Natur.400..242O.  doi:10.1038/22281. 
S2CID  4399615.
65. ^ Jump up to:a b 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".  Physical Review C. 69 (5):
054607. Bibcode:2004PhRvC..69e4607O. doi:10.1103/P
hysRevC.69.054607.
66. ^ Barber, R. C.; Gäggeler, H. W.; Karol, P. J.; Nakahara,
H.; Vardaci, E.; Vogt, E. (2009).  "Discovery of the
element with atomic number 112 (IUPAC Technical
Report)"  (PDF). Pure and Applied Chemistry. 81 (7):
1331.  doi:10.1351/PAC-REP-08-03-05. S2CID  9570383
3.
67. ^ Barber, R. C.; Karol, P. J.; Nakahara, H.; Vardaci, E.;
Vogt, E. W. (2011).  "Discovery of the elements with
atomic numbers greater than or equal to 113 (IUPAC
Technical Report)".  Pure and Applied Chemistry.  83  (7):
1485.  doi:10.1351/PAC-REP-10-05-01.
68. ^ 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, Alexander (9
July 2016).  "A new assessment of the alleged link
between element 115 and element 117 decay
chains"(PDF).  Physics Letters B.  760  (2016): 293–
6. Bibcode:2016PhLB..760..293F.  doi:10.1016/j.physletb
.2016.07.008. Retrieved 2 April 2016.
69. ^ Forsberg, Ulrika; Fahlander, Claes; Rudolph, Dirk
(2016).  Congruence of decay chains of elements 113,
115, and 117  (PDF). Nobel Symposium NS160 –
Chemistry and Physics of Heavy and Superheavy
Elements. doi:10.1051/epjconf/201613102003.
70. ^ Morita, Kōsuke (2014).  "Research on Superheavy
Elements at RIKEN"  (PDF).  APS Division of Nuclear
Physics Meeting Abstracts. 2014:
DG.002.  Bibcode:2014APS..DNP.DG002M.
Retrieved 28 April  2017.
71. ^ Morimoto, Kouji (October 2009).  "Production and
Decay Properties of  266Bh and its daughter nuclei by
using the  248Cm(23Na,5n)266Bh
Reaction"  (PDF). www.kernchemie.uni-mainz.de.  Univer
sity of Mainz. Archived from  the original  (PDF)  on 21
September 2017. Retrieved  28 April 2017.
72. ^ Jump up to:a b c d e f g h i Zagrebaev, Valeriy; Karpov,
Alexander; Greiner, Walter (2013). "Future of
superheavy element research: Which nuclei could be
synthesized within the next few years?"  (PDF).  Journal of
Physics: Conference Series. 420. IOP Science. pp. 1–
15. Retrieved  20 August 2013.
73. ^ Heinz, Sophie (1 April 2015). "Probing the Stability of
Superheavy Nuclei with Radioactive Ion
Beams"  (PDF). cyclotron.tamu.edu. Texas A & M
University. Retrieved 30 April  2017.
74. ^ Chatt, J. (1979). "Recommendations for the naming of
elements of atomic numbers greater than 100". Pure and
Applied Chemistry. 51 (2): 381–
384.  doi:10.1351/pac197951020381.
75. ^ Koppenol, W. H. (2002). "Naming of new elements
(IUPAC Recommendations 2002)"(PDF).  Pure and
Applied Chemistry. 74 (5):
787.  doi:10.1351/pac200274050787. S2CID  95859397.
76. ^ Brown, M. (6 June 2011).  "Two Ultraheavy Elements
Added to Periodic Table".  Wired. Retrieved  7 June  2011.
77. ^ Jump up to:a b Welsh, J. (2 December 2011).  "Two
Elements Named: Livermorium and
Flerovium". LiveScience. Retrieved 2 December 2011.
78. ^ "Российские физики предложат назвать 116
химический элемент московием"[Russian physicists
have offered to call 116 chemical
element  moscovium].  RIA Novosti. 26 March 2011.
Retrieved 8 May  2011. Mikhail Itkis, the vice-director of
JINR stated: "We would like to name element 114
after Georgy Flerov – flerovium, and the second [element
116] – moscovium, not after Moscow, but after Moscow
Oblast".
79. ^ Popeko, Andrey G. (2016). "Synthesis of superheavy
elements"  (PDF). jinr.ru.  Joint Institute for Nuclear
Research. Retrieved  4 February  2018.
80. ^ Oganessian, Yu. Ts. (10 October 2015).  "Гамбургский
счет" [Hamburg reckoning] (Interview) (in Russian).
Interviewed by Orlova, O.  Public Television of Russia.
Retrieved 18 January  2020.
81. ^ Kalinkin, B. N.; Gareev, F. A. (2001).  Synthesis of
Superheavy elements and Theory of Atomic
Nucleus. Exotic Nuclei. p.  118. arXiv:nucl-
th/0111083v2. Bibcode:2002exnu.conf..118K. CiteSeerX 
10.1.1.264.7426. doi:10.1142/9789812777300_0009.  IS
BN  978-981-238-025-8. S2CID  119481840.
82. ^ Jump up to:a b "JINR Annual Reports 2000–2006". JINR.
Retrieved 27 August  2013.
83. ^ Jump up to:a b Emsley, John (2011).  Nature's Building
Blocks: An A-Z Guide to the Elements (New ed.). New
York, NY: Oxford University Press. p.  590. ISBN 978-0-
19-960563-7.
84. ^ Jump up to:a b c Zagrebaev, V.; Greiner, W. (2008).
"Synthesis of superheavy nuclei: A search for new
production reactions".  Physical Review C. 78 (3):
034610. arXiv:0807.2537. Bibcode:2008PhRvC..78c461
0Z. doi:10.1103/PhysRevC.78.034610.
85. ^ Chowdhury, P. R.; Samanta, C.; Basu, D. N. (2006). "α
decay half-lives of new superheavy elements".  Physical
Review C.  73  (1): 014612.  arXiv:nucl-
th/0507054.  Bibcode:2006PhRvC..73a4612C. doi:10.11
03/PhysRevC.73.014612.  S2CID 118739116.
86. ^ Samanta, C.; Chowdhury, P. R.; Basu, D. N. (2007).
"Predictions of alpha decay half lives of heavy and
superheavy elements". Nuclear Physics A. 789 (1–4):
142–154. arXiv:nucl-th/0703086.  Bibcod
e:2007NuPhA.789..142S.  CiteSeerX 10.1.1.264.8177. d
oi:10.1016/j.nuclphysa.2007.04.001.  S2CID 7496348.
87. ^ Chowdhury, P. R.; Samanta, C.; Basu, D. N. (2008).
"Search for long lived heaviest nuclei beyond the valley
of stability". Physical Review C.  77  (4):
044603. arXiv:0802.3837. Bibcode:2008PhRvC..77d460
3C.  doi:10.1103/PhysRevC.77.044603. S2CID  1192078
07.
88. ^ Roy Chowdhury, P.; Samanta, C.; Basu, D. N. (2008).
"Nuclear half-lives for α-radioactivity of elements with 100
≤ Z ≤ 130". Atomic Data and Nuclear Data
Tables. 94 (6): 781–
806.  arXiv:0802.4161.  Bibcode:2008ADNDT..94..781C. 
doi:10.1016/j.adt.2008.01.003.
89. ^ Thayer 2010, pp. 63–64.
90. ^ Faegri, K.; Saue, T. (2001). "Diatomic molecules
between very heavy elements of group 13 and group 17:
A study of relativistic effects on bonding".  Journal of
Chemical Physics.  115  (6):
2456.  Bibcode:2001JChPh.115.2456F.  doi:10.1063/1.13
85366.
91. ^ Thayer 2010, pp. 63–67.
92. ^ Borschevsky, Anastasia; Pershina, Valeria; Kaldor, Uzi;
Eliav, Ephraim.  "Fully relativistic  ab initio  studies of
superheavy elements"  (PDF).  www.kernchemie.uni-
mainz.de.  Johannes Gutenberg University Mainz.
Archived from  the original  (PDF)  on 15 January 2018.
Retrieved 15 January  2018.
93. ^ Hermann, Andreas; Furthmüller, Jürgen; Gäggeler,
Heinz W.; Schwerdtfeger, Peter (2010).  "Spin-orbit
effects in structural and electronic properties for the solid
state of the group-14 elements from carbon to
superheavy element 114".  Physical Review B.  82  (15):
155116–1–8.  Bibcode:2010PhRvB..82o5116H.  do
i:10.1103/PhysRevB.82.155116.
94. ^ Jump up to:a b Thayer 2010, pp. 64.
95. ^ Zaitsevskii, A.; van Wüllen, C.; Rusakov, A.; Titov, A.
(September 2007). "Relativistic DFT and ab initio
calculations on the seventh-row superheavy elements:
E113 - E114"  (PDF).  jinr.ru. Retrieved 17
February  2018.
96. ^ Jump up to:a b Pershina 2010, p. 502.
97. ^ Pershina 2010, p. 503.
98. ^ Jump up to:    Thayer 2010, p. 83.
a b

99. ^ Fricke, B.; Greiner, W.; Waber, J. T. (1971). "The


continuation of the periodic table up to Z = 172. The
chemistry of superheavy elements"  (PDF). Theoretica
Chimica Acta.  21  (3): 235–
260.  doi:10.1007/BF01172015. S2CID  117157377.
100. ^ Balasubramanian, K. (30 July 2002).
"Breakdown of the singlet and triplet nature of electronic
states of the superheavy element 114 dihydride
(114H2)".  Journal of Chemical Physics.  117  (16): 7426–
32.  Bibcode:2002JChPh.117.7426B. doi:10.1063/1.1508
371.
101. ^ Winter, M. (2012). "Flerovium: The
Essentials".  WebElements. University of Sheffield.
Retrieved 28 August  2008.
102. ^ Jump up to:a b c d "Flerov Laboratory of Nuclear
Reactions"  (PDF). 2009. pp. 86–96. Retrieved 1
June  2012.
103. ^ Jump up to:a b c d Eichler, Robert; Aksenov, N. V.;
Albin, Yu. V.; Belozerov, A. V.; Bozhikov, G. A.;
Chepigin, V. I.; Dmitriev, S. N.; Dressler, R.; Gäggeler, H.
W.; Gorshkov, V. A.; Henderson, G. S.
(2010).  "Indication for a volatile element
114"  (PDF). Radiochimica Acta.  98  (3): 133–
139.  doi:10.1524/ract.2010.1705. S2CID  95172228.
104. ^ Kratz, Jens Volker (2012). "The impact of the
properties of the heaviest elements on the chemical and
physical sciences". Radiochimica Acta.  100  (8–9): 569–
578.  doi:10.1524/ract.2012.1963. S2CID  97915854.
105. ^ Düllmann, Christoph E. (18 September
2012). Superheavy element 114 is a volatile metal.
Archived from  the original on 27 September 2013.
Retrieved 25 September2013.
106. ^ Yakushev, Alexander; Gates, Jacklyn M.; Türler,
Andreas; Schädel, Matthias; Düllmann, Christoph E.;
Ackermann, Dieter; Andersson, Lise-Lotte; Block,
Michael; Brüchle, Willy; Dvorak, Jan; Eberhardt, Klaus;
Essel, Hans G.; Even, Julia; Forsberg, Ulrika; Gorshkov,
Alexander; Graeger, Reimar; Gregorich, Kenneth E.;
Hartmann, Willi; Herzberg, Rolf-Deitmar; Heßberger,
Fritz P.; Hild, Daniel; Hübner, Annett; Jäger, Egon;
Khuyagbaatar, Jadambaa; Kindler, Birgit; Kratz, Jens V.;
Krier, Jörg; Kurz, Nikolaus; Lommel, Bettina; Niewisch,
Lorenz J.; Nitsche, Heino; Omtvedt, Jon Petter; Parr,
Edward; Qin, Zhi; Rudolph, Dirk; Runke, Jörg;
Schausten, Birgitta; Schimpf, Erwin; Semchenkov,
Andrey; Steiner, Jutta; Thörle-Pospiech, Petra; Uusitalo,
Juha; Wegrzecki, Maciej; Wiehl, Norbert
(2014).  "Superheavy Element Flerovium (Element 114)
Is a Volatile Metal"  (PDF). Inorg. Chem. 53 (1624):
1624–1629.  doi:10.1021/ic4026766.  PMID  24456007.
Retrieved 30 March2017.
107. ^ Yakushev, Alexander; Eichler, Robert
(2016).  Gas-phase chemistry of element 114,
flerovium  (PDF). Nobel Symposium NS160 – Chemistry
and Physics of Heavy and Superheavy
Elements. doi:10.1051/epjconf/201613107003.
108. ^ Moody, Ken (30 November 2013). "Synthesis of
Superheavy Elements". In Schädel, Matthias;
Shaughnessy, Dawn (eds.).  The Chemistry of
Superheavy Elements (2nd ed.). Springer Science &
Business Media. pp. 24–8.  ISBN  9783642374661.

Bibliography[edit]
 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. doi:1
0.1088/1674-1137/41/3/030001.
 Beiser, A. (2003). Concepts of modern
physics (6th ed.). McGraw-Hill. ISBN 978-0-07-
244848-1. OCLC 48965418.
 Hoffman, D. C.; Ghiorso, A.; Seaborg, G. T.
(2000). The Transuranium People: The Inside
Story. World Scientific. ISBN 978-1-78-326244-
1.
 Kragh, H. (2018). From Transuranic to
Superheavy Elements: A Story of Dispute and
Creation. Springer. ISBN 978-3-319-75813-8.
 Zagrebaev, 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. Bibcode:2013JPhCS.
420a2001Z. doi:10.1088/1742-
6596/420/1/012001. ISSN 1742-6588. S2CID 5
5434734.

Bibliography[edit]
 Barysz, M.; Ishikawa, Y., eds. (2010). Relativisic Methods for
Chemists. Springer. ISBN 978-1-4020-9974-8.

 Thayer, J. S. (2010). "Relativistic Effects and the


Chemistry of the Heavier Main Group
Elements".  Relativistic Methods for Chemists. Challenges
and Advances in Computational Chemistry and
Physics. 10. pp. 63–97. doi:10.1007/978-1-4020-9975-
5_2.  ISBN  978-1-4020-9974-8.
 Stysziński, J. (2010). Why do we need relativistic
computational methods?. p. 99.
 Pershina, V. (2010).  Electronic structure and chemistry of
the heaviest elements. p. 450.

External links[edit]
 CERN Courier – First postcard from the
island of nuclear stability
 CERN Courier – Second postcard from the
island of stability
hide
Periodic table (Large cells)
4 5 6 7 8 9 10 11 12 13 14

B C
Al Si
Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge
Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn
Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Hf Ta W Re Os Ir Pt Au Hg Tl Pb
Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr Rf Db Sg Bh Hs Mt Ds Rg Cn Nh Fl

earth metal Lanthanide Actinide Transition metal Post-transition metal Metalloid Reactive nonmetal Nobl

LCCN: sh2012003542

Categories: 
 Flerovium
 Chemical elements
 Synthetic elements
Navigation menu
 Not logged in
 Talk
 Contributions
 Create account
 Log in
 Article
 Talk
 Read
 Edit
 View history
Search
Search Go

 Main page
 Contents
 Current events
 Random article
 About Wikipedia
 Contact us
 Donate
Contribute
 Help
 Learn to edit
 Community portal
 Recent changes
 Upload file
Tools
 What links here
 Related changes
 Special pages
 Permanent link
 Page information
 Cite this page
 Wikidata item
Print/export
 Download as PDF
 Printable version
In other projects
 Wikimedia Commons
Languages
 ‫العربية‬
 Cebuano
 Español
 Bahasa Indonesia
 Bahasa Melayu
 Русский
 Tagalog
 Winaray
 中文
103 more
Edit links
 This page was last edited on 21 October 2020, at 17:23 (UTC).
 Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply.
By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark
of the Wikimedia Foundation, Inc., a non-profit organization.
 Privacy policy

 About Wikipedia

 Disclaimers

 Contact Wikipedia

 Mobile view

 Developers

 Statistics

 Cookie statement

You might also like