You are on page 1of 3

Livermorium is a synthetic chemical element with the symbol Lv and has an atomic

number of 116. It is an extremely radioactive element that has only been created in
a laboratory setting and has not been observed in nature. The element is named
after the Lawrence Livermore National Laboratory in the United States, which
collaborated with the Joint Institute for Nuclear Research (JINR) in Dubna, Russia,
to discover livermorium during experiments conducted between 2000 and 2006. The
name of the laboratory refers to the city of Livermore, California, where it is
located, which in turn was named after the rancher and landowner Robert Livermore.
The name was adopted by IUPAC on May 30, 2012.[6] Four isotopes of livermorium are
known, with mass numbers between 290 and 293 inclusive; the longest-lived among
them is livermorium-293 with a half-life of about 60 milliseconds. A fifth possible
isotope with mass number 294 has been reported but not yet confirmed.

In the periodic table, it is a p-block transactinide element. It is a member of the


7th period and is placed in group 16 as the heaviest chalcogen, but it has not been
confirmed to behave as the heavier homologue to the chalcogen polonium. Livermorium
is calculated to have some similar properties to its lighter homologues (oxygen,
sulfur, selenium, tellurium, and polonium), and be a post-transition metal, though
it should also show several major differences from them.

Introduction
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
A graphic depiction of a nuclear fusion reaction. Two nuclei fuse into one,
emitting a neutron. Thus far, reactions that created new elements were similar,
with the only possible difference that several singular neutrons sometimes were
released, or none at all.
External video
video icon Visualization of unsuccessful nuclear fusion, based on calculations by
the Australian National University[8]
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.[14] 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 fuse into one only 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.[15] 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.[15][16] 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.[17][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.[20] 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.[20] The transfer
takes about 10−6 seconds; in order to be detected, the nucleus must survive this
long.[23] The nucleus is recorded again once its decay is registered, and the
location, the energy, and the time of the decay are measured.[20]
Stability of a nucleus is provided by the strong interaction. However, its range is
very short; as nuclei become larger, their 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.[24] Nuclei of
the heaviest elements are thus theoretically predicted[25] and have so far been
observed[26] 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
Unsuccessful synthesis attempts
The first search for element 116, using the reaction between 248Cm and 48Ca, was
performed in 1977 by Ken Hulet and his team at the Lawrence Livermore National
Laboratory (LLNL). They were unable to detect any atoms of livermorium.[38] Yuri
Oganessian and his team at the Flerov Laboratory of Nuclear Reactions (FLNR) in the
Joint Institute for Nuclear Research (JINR) subsequently attempted the reaction in
1978 and met failure. In 1985, in a joint experiment between Berkeley and Peter
Armbruster's team at GSI, the result was again negative, with a calculated cross
section limit of 10–100 pb. Work on reactions with 48Ca, which had proved very
useful in the synthesis of nobelium from the natPb+48Ca reaction, nevertheless
continued at Dubna, with a superheavy element separator being developed in 1989, a
search for target materials and starting of collaborations with LLNL being started
in 1990, production of more intense 48Ca beams being started in 1996, and
preparations for long-term experiments with 3 orders of magnitude higher
sensitivity being performed in the early 1990s. This work led directly to the
production of new isotopes of elements 112 to 118 in the reactions of 48Ca with
actinide targets and the discovery of the 5 heaviest elements on the periodic
table: flerovium, moscovium, livermorium, tennessine, and oganesson.[39]

In 1995, an international team led by Sigurd Hofmann at the Gesellschaft für


Schwerionenforschung (GSI) in Darmstadt, Germany attempted to synthesise element
116 in a radiative capture reaction (in which the compound nucleus de-excites
through pure gamma emission without evaporating neutrons) between a lead-208 target
and selenium-82 projectiles. No atoms of element 116 were identified.[40]

Unconfirmed discovery claims


In late 1998, Polish physicist Robert Smolańczuk published calculations on the
fusion of atomic nuclei towards the synthesis of superheavy atoms, including
elements 118 and 116.[41] His calculations suggested that it might be possible to
make these two elements by fusing lead with krypton under carefully controlled
conditions.[41]

In 1999, researchers at Lawrence Berkeley National Laboratory made use of these


predictions and announced the discovery of elements 118 and 116, in a paper
published in Physical Review Letters,[42] and very soon after the results were
reported in Science.[43] The researchers reported to have performed the reaction

86
36Kr
+ 208
82Pb
→ 293
118Og
+
n
→ 289
116Lv
+ α
The following year, they published a retraction after researchers at other
laboratories were unable to duplicate the results and the Berkeley lab itself was
unable to duplicate them as well.[44] In June 2002, the director of the lab
announced that the original claim of the discovery of these two elements had been
based on data fabricated by principal author Victor Ninov.[45][46]

Discovery
Livermorium was first synthesized on July 19, 2000, when scientists at Dubna (JINR)
bombarded a curium-248 target with accelerated calcium-48 ions. A single atom was
detected, decaying by alpha emission with decay energy 10.54 MeV to an isotope of
flerovium. The results were published in December 2000.[47]

248
96Cm
+ 48
20Ca
→ 296
116Lv
* → 293
116Lv
+ 3 1
0n
→ 289
114Fl
+ α
The daughter flerovium isotope had properties matching those of a flerovium isotope
first synthesized in June 1999, which was originally assigned to 288Fl,[47]
implying an assignment of the parent livermorium isotope to 292Lv. Later work in
December 2002 indicated that the synthesized flerovium isotope was actually 289Fl,
and hence the assignment of the synthesized livermorium atom was correspondingly
altered to 293Lv.[48]

You might also like